You are on page 1of 14

Applied Catalysis A: General 439–440 (2012) 111–124

Contents lists available at SciVerse ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Hydrodeoxygenation of guaiacol over carbon-supported molybdenum nitride


catalysts: Effects of nitriding methods and support properties
I. Tyrone Ghampson a,b , Catherine Sepúlveda c , Rafael Garcia c , Ljubisa R. Radovic d,e , J.L. García Fierro f ,
William J. DeSisto a,g,∗ , Nestor Escalona c,∗∗
a
Department of Chemical and Biological Engineering, University of Maine, Orono, ME 04469, United States
b
Unidad de Desarrollo Tecnológico, Universidad de Concepcićn, Casilla 4051, Concepcićn, Chile
c
Universidad de Concepćion, Facultad de Ciencias Quimicas, Casilla 160c, Concepcićn, Chile
d
Penn State University, University Park, PA 16802, United States
e
Universidad de Concepcićn, Facultad de Ingenieria, Dept. Ing. Quimica, Concepcićn, Chile
f
Instituto de Catalisis y Petroquimica, CSIC, Cantoblanco, 28049 Madrid, Spain
g
Forest Bioproducts Research Institute, University of Maine, Orono, ME 04469, United States

a r t i c l e i n f o a b s t r a c t

Article history: Molybdenum nitride catalysts supported on activated carbon materials with different textural and chem-
Received 25 April 2012 ical properties were synthesized by nitriding supported Mo oxide precursors with gaseous NH3 or N2 /H2
Received in revised form 25 June 2012 mixtures using a temperature-programmed reaction. The supports and catalysts were characterized by
Accepted 28 June 2012
N2 physisorption, XRD, chemical analysis, TPD, FT-IR and XPS. Guaiacol (2-methoxyphenol) hydrodeoxy-
Available online 5 July 2012
genation (HDO) activities at 5 MPa and 300 ◦ C were evaluated in a batch autoclave reactor. Molybdenum
nitrides prepared using a N2 /H2 mixture resulted in more highly dispersed catalysts, and consequently
Keywords:
more active catalysts, relative to those prepared using ammonolysis. The HDO activity was also related to
Hydrodeoxygenation
Guaiacol
pore size distribution and the concentration of oxygen-containing surface groups of the different carbon
Activated carbon supports. Increased mesoporosity is argued to have facilitated the access to active sites while increased
Mo2 N catalysts surface acidity enhanced their catalytic activity through modification of their electronic properties. The
highest activity was thus attributed to the highest dispersion of the unsaturated catalyst species and the
highest support mesoporosity. Surprisingly, addition of Co did not improve the HDO activity.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction prevention of coke formation/catalyst deactivation and (ii) selec-


tive removal of oxygen without excessive hydrogenation of
Due to long-term economic and environmental concerns, bio-oil aromatic and olefinic compounds [2,3].
derived from pyrolysis of woody biomass has received consider- Model compounds have been used to mimic HDO studies of
able attention as an alternative renewable feedstock to crude oil bio-oil components in an effort to understand the role and fate of
for the production of fuels and value-added chemicals [1]. Its uti- different functional groups present in the feed, as well as provide
lization as fuel is limited, however, by high viscosity, low heating additional insight into the development of improved catalysts and
value, incomplete volatility and thermal instability, which stem processes [3]. Guaiacol (2-methoxyphenol) is commonly used as a
from the relative abundance of oxygenated organic compounds model compound for HDO studies to represent the large number
[2]. Catalytic hydrodeoxygenation (HDO) reactions are typically of mono- and dimethoxy phenols present in bio-oil [4]; it is known
performed to refine bio-oil and increase its quality as transporta- to be a precursor to catechol but it also subsequently forms coke
tion fuel. There are two significant challenges in this process: (i) [5,6]. Also, guaiacol possesses two different oxygenated functional
groups ( OCH3 and OH) which make it challenging to achieve
complete deoxygenation [7].
Heterogeneous catalysts commonly studied for HDO of guaia-
∗ Corresponding author at: Department of Chemical and Biological Engineering, col (and many other model compounds) are conventional sulfided
University of Maine, Orono, ME 04469, United States. Tel.: +1 207 581 2291; Co(Ni)Mo/␥-Al2 O3 [5,8] and supported noble metal catalysts such
fax: +1 207 581 2323.
∗∗ Corresponding author at: Tel.: +56 41 2207236; fax: +56 41 2245374. as Ru, Rh and Pd [9,10]. The initial interest in the metal sulfides
E-mail addresses: WDeSisto@umche.maine.edu (W.J. DeSisto), was driven by high cost and lack of selective HDO activity of
nescalona@udec.cl (N. Escalona). the noble metal catalysts. Despite the high catalytic activity for

0926-860X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2012.06.047
112 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

guaiacol conversion, there are some drawbacks associated with ammonolysis or reduction/nitridation using a N2 /H2 mixture.
these sulfide catalysts: (i) The alumina support can be unstable in The effects of the synthesis procedure, support properties and the
water at processing conditions. (ii) The sulfide catalyst can oxi- addition of Co as a promoter on the HDO of guaiacol were examined
dize under processing conditions, requiring in situ regeneration in terms of catalytic activity and phenol/catechol selectivity.
with a sulfiding agent to prolong catalyst activity; this regenera-
tion can contaminate products [5,6,11,12]. (iii) The acidic nature 2. Experimental
of the alumina support is known to be the cause of substantial
coke deposition and rapid catalyst deactivation [13]. These draw- 2.1. Preparation of catalysts
backs prompted interest in less acidic materials such as silica [14],
zirconia [9,15] and activated carbon [16,17] as catalyst supports. Four commercial activated carbons (Norit Americas, Inc.) were
Centeno et al. [14] reported that despite the lower activity of used as supports: NORIT GAC 1240 Plus (0.42–2.00 mm particle
metal sulfides supported on silica and carbon compared with con- size, SBET = 976 m2 g−1 , total pore volume = 0.56 cm3 g−1 ), NORIT
ventional alumina-supported counterparts, the use of alternative GCA 1240 Plus (0.42–1.70 mm, 1132 m2 g−1 , 0.51 cm3 g−1 ), Darco
supports led to negligible coke formation. Furthermore, studies MRX (0.60–2.00 mm, 613 m2 g−1 , 0.62 cm3 g−1 ), and NORIT CGran
involving carbon- and zirconia-supported catalysts indicated direct (0.50–1.70 mm, 1402 m2 g−1 , 1.15 cm3 g−1 ). Prior to their use, the
elimination of the methoxy group which favored direct production activated carbon materials were treated with 1 M HNO3 at 90 ◦ C
of phenol from guaiacol [14,18]. Particularly, based on their supe- for 6 h. The solution was then filtered and extensively washed with
rior performance in hydrodesulfurization processes, carbons are distilled water to bring the pH to ca. 7. The samples were dried
known to be promising supports for the HDO of bio-oil [19–21]. overnight under vacuum at 120 ◦ C. The supported molybdenum
Interest in carbon supports has increased mainly due to its oxide precursors were prepared by incipient wetness impregnation
remarkable flexibility and the ability to recover active metal after using aqueous solutions of ammonium heptamolybdate (Fischer,
catalyst deactivation [20]. For HDO reactions, such deactivation in AHM, (NH4 )6 Mo7 O24 ·4H2 O, A.C.S. grade). After the impregnation,
the presence of water could be limited due to the hydrophobic char- the samples were kept at room temperature for 24 h, followed
acter of the carbon surface [22]. On the other hand, the weaker by drying overnight at 110 ◦ C. The bimetallic oxide precursors
interaction between the support surface and the active metal may were prepared by sequential impregnation: Mo-loaded samples
result in a lower dispersion of the sulfide phase [14,20]. The car- were first prepared using the same drying–calcination procedure
bon surface can be ‘decorated’ with oxygen functionalities; this described above; these samples were then impregnated with aque-
improves catalytic activity by facilitating a higher dispersion of the ous solution of cobalt (II) nitrate hexahydrate (Acros Organics,
active phase [23,24]. For example, oxidative treatments with HNO3 Co(NO3 )2 ·6H2 O, 99%) and kept overnight at room temperature fol-
modified the carbon surface chemistry and promoted the formation lowed by drying overnight at 110 ◦ C. The supported oxides were
of small, well-dispersed crystals of the molybdenum precursor on prepared to obtain a nominal loading of 10 wt% Mo for monometal-
the support [21,25] although this led to lower phenol yields during lic samples and 10 wt% Mo plus 2.4 wt% Co for the bimetallic
guaiacol HDO [23]. Additional studies further confirmed that HDO samples. All oxide precursors were sieved to obtain 180–450 ␮m
chemistry can be controlled by modifying the surface chemistry of particle size range.
the carbon support and consequently the dispersion of the catalyst Molybdenum nitrides were prepared by loading a 10 mm i.d.
[16,22]. This adds to the potential use of carbon-supported systems quartz reactor tube with 2.5 g of the oxidic precursor, while pass-
in rational catalyst design [17,19]. ing ammonia (Matheson Tri-Gas, NH3 , 99.99%) or a N2 /H2 mixture
A wide variety of active phases – metals [4,9,10,26], transition (N2 : BOC Gases, Grade 5; H2 : Matheson Tri-Gas, 99.99%) over the
metal phosphides [7] and transition metal nitrides [27] – have been sample [33,34]. The reactor was initially purged with nitrogen for
employed for HDO reactions in order to obviate the need to add sul- 30 min and switched to NH3 (300 mL min−1 ) or a N2 /H2 mixture
fur to the feed. In particular, transition metal nitrides show great (300 mL min−1 , N2 /H2 = 5/1 (v/v)). The temperature was linearly
potential as catalysts due to their ceramic-like physical properties increased from ambient temperature to 300 ◦ C within 30 min
coupled with chemical properties resembling platinum-group met- (9.33 ◦ C min−1 ), then from 300 to 500 ◦ C at 0.6 ◦ C min−1 , and from
als [28]. These materials are also responsible for unique catalytic 500 to 700 ◦ C at 2 ◦ C min−1 . The temperature was maintained at
pathways, leading to desirable product selectivities [28,29]. Con- 700 ◦ C for 2 h. The nitrides prepared using NH3 were cooled to room
sequently, they offer a cheaper and more selective alternative to temperature using the same flow rate of NH3 while the nitrides
noble-metal catalysts such as Ru, Pd and Pt. Our previous study prepared using the N2 /H2 mixture were cooled in 300 mL min−1 of
showed high activities and a high phenol/catechol ratio for bulk N2 . The materials were then passivated in 1% O2 /N2 (BOC Gases,
molybdenum nitride catalysts in the HDO of guaiacol [30]. How- UHP grade) for 12 h to avoid oxidation upon exposure to air. The
ever, supported catalysts are preferred in commercial applications preparation condition was adopted after careful review of the lit-
for mechanical and morphological stability. The addition of Co as a erature in addition to previous results to ascertain the optimal
promoter is known to improve the activity of bulk and supported condition for unsupported molybdenum nitrides [30]. Preparation
Mo2 N catalysts for HDS and HDN reactions [31,32]. Furthermore, of nitride catalysts using NH3 and N2 /H2 is referred to as method
Co-promoted MoS2 catalysts have been reported to exhibit sig- 1 and method 2, respectively. For notation, Mo nitrides prepared
nificantly higher HDO activity compared to non-promoted MoS2 using method 1 have suffix “A”, while suffix “NH” implies method 2:
catalysts for HDO of guaiacol [8]. Although addition of Co improved e.g., MoN/Darco-A and MoN/Darco-NH are Darco activated carbon-
the yield of deoxygenated products, the overall activity was not supported Mo nitride catalysts prepared using NH3 and N2 /H2 ,
enhanced compared to the monometallic nitride [30]. This moti- respectively.
vated us to investigate also the effect of Co promoters on the Molybdenum, cobalt and nitrogen contents in the catalysts were
catalytic properties of supported nitrides in HDO reactions. performed by Galbraith Laboratories using ICP-AES for the metals
Here we report on the behavior of molybdenum nitrides and a combustion method for nitrogen.
dispersed on four different activated carbon supports. The
supports were both microporous/mesoporous and meso- 2.2. Nitrogen porosimetry
porous/macroporous carbons. The catalysts were synthesized
by impregnation of an aqueous salt and its subsequent conver- Nitrogen adsorption/desorption isotherms were obtained at
sion to the nitride; thermal conversion was achieved by either 77 K using a Micromeritics ASAP-2020 instrument to evaluate the
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 113

porous structure of the support and catalyst samples. Prior to sample was first pretreated at 100 ◦ C for 4 h in He (50 mL min−1 ) to
the measurements, the samples were outgassed under vacuum remove most of the weakly adsorbed water, and cooled to room
at 200 ◦ C for 12 h. The isotherms were collected within a broad temperature in He. The pretreated sample was then heated in
relative pressure range, 10−6 < P/P0 < 0.995, and using a low pres- a flow of He (50 mL min−1 ) from room temperature to 800 ◦ C at
sure incremental dosing (3 cm3 g−1 STP) in order to obtain an 10 ◦ C min−1 .
adequate characterization of the micropore region. The isotherms
were used to calculate the BET specific surface area (SBET ), total 2.6. Fourier transform-infrared transmission (FT-IR) spectroscopy
pore volume (TPV), average pore diameter (dpore ), and micropore
volume (V ). SBET was obtained from the adsorption branch in FT-IR analyses of the supports were performed on a Nicolet
the range 0.04 ≤ P/P0 ≤ 0.14 and TPV was recorded at P/P0 = 0.995. Nexus FTIR in the wavenumber range (4000–400 cm−1 ) and with a
Average pore diameters were calculated from the equation dpore = scan of 64. The samples were prepared using a 1:100 mg of activated
2 · TPV/SBET , assuming slit-shaped pores. The pore size distribu- carbon and KBr support.
tions (PSD) for 0.4–100 nm were determined from the adsorption
branch of the isotherm using the nonlocal density functional theory 2.7. Acidity measurements
(NLDFT) method [35,36]. Micropore volume was calculated from
the NLDFT cumulative volume of pores whose size was below 2 nm. Acid site concentrations and acid strength measurements of the
supports and selected catalysts were determined using a potentio-
2.3. X-ray diffraction metric method [38], whereby a suspension in acetonitrile (Merck,
99.9%) was titrated with n-butylamine (Merck, 99%). The varia-
Wide-angle X-ray diffraction (XRD) patterns of powdered sam- tion in electric potential was registered on a Denver Instrument
ples were obtained using a PANalytical X’Pert Pro diffractometer UltraBasic pH/mV meter.
equipped with a graphite monochromator and Cu K␣ radiation
(45 kV, 40 mA) in a parallel beam optical geometry. The standard 2.8. Reaction characterization
scan parameters were 15–85◦ 2 with a step size of 0.02◦ and a
counting time of 10 s per step. Identification of the phases was Reactivity studies were performed in a 300 mL stirred-batch
achieved by reference to the JCPDS data files. autoclave set-up (Parr Model 4841) at 300 ◦ C and under a hydrogen
pressure of 5 MPa. Prior to catalyst testing, the passivated samples
2.4. X-ray photoelectron spectroscopy (XPS) were activated ex situ under H2 (AGA Chile, 99.99%) at a flow rate
of 60 mL min−1 and 450 ◦ C for 6 h, conditions that were shown in
X-ray photoelectron spectra of reduced catalysts were obtained previous studies to maximize HDO conversion [27]. Approximately
on a VG Escalab 200R electron spectrometer using a Mg K␣ 0.25 g of freshly pre-treated catalyst was added to the reactor
(1253.6 eV) photon source. The samples were pre-reduced ex situ charged with 80 mL of decalin (Merck, 99%), 2.53 mL of guaiacol
with H2 at 450 ◦ C for 6 h. After reduction, the samples were cooled (0.232 mol L−1 , Merck, 99.5%), and 700 ␮L of hexadecane (Merck,
to room temperature, flushed with nitrogen and stored in flasks 99%). Hexadecane was used as an internal standard for quanti-
containing isooctane (Merck, 99.8%), then transferred to the pre- tative GC analysis. The sealed reactor was flushed with nitrogen
treatment chamber of the spectrometer. The binding energies (BE) (AGA Chile, Grade 5) for 30 min to evacuate air from the system.
were referenced to the C 1s level of the carbon support at 284.9 eV. While continuously stirring the mixture, the reactor was heated to
An estimated error of ±0.1 eV can be assumed for all measurements. 300 ◦ C under N2 . Once the reaction temperature was attained, N2
Intensities of the peaks were calculated from the respective peak was replaced with H2 and then pressurized to 5 MPa. This pressure
areas after background subtraction and spectrum fitting by a com- was maintained for the entire duration of the experiment by adding
bination of Gaussian/Lorentzian functions. Relative surface atomic H2 to the reactor whenever necessary. Liquid samples were peri-
ratios were calculated from odically withdrawn during the course of the reaction after purging
the sampling line by withdrawing a small amount of the reactant
(SMo 3d /fMo 3d ) mixture. The samples were analyzed by a Perkin Elmer (Clarus
(Mo/C)atomic ratio = ; (Co/C)atomic ratio
(SC 1s /fC 1s ) 400) gas chromatograph equipped with a flame ionization detector
(SCo 2p /fCo 2p ) (SN 1s /fN 1s ) (FID) and a CP-Sil 5 CB column (Agilent, 30 m × 0.53 mm × 1.0 ␮m
= ; (N/C)atomic ratio = film thickness). The injector and FID were held at 275 and 180 ◦ C,
(SC 1s /fC 1s ) (SC 1s /fC 1s )
respectively. (The GC oven program consisted of an initial isother-
where S is the corresponding peak areas and f is the respective mal operation at 30 ◦ C for 6 min, followed by heating to 70 ◦ C at
tabulated sensitivity factor [37] and the factor for the device. The 30 ◦ C min−1 with an isotherm of 22 min, and a subsequent ramp
atomic ratios were calculated at a precision of 7%. to 275 ◦ C at 30 ◦ C min−1 .) The product distributions were identified
by their column retention time in comparison with available stan-
2.5. Temperature-programmed decomposition–mass dards. The initial concentration of guaiacol was taken as 100% in
spectroscopy (TPD–MS) order to ignore slight conversion before isothermal condition was
achieved. The catalytic activity was expressed as the initial reaction
TPD analyses of the carbon supports were carried out in an rate, calculated from the slope of the conversion vs. time plot, as
in-house apparatus which consisted of a U-shaped quartz tube well as by the intrinsic activity (in moles of guaiacol consumed per
micro-reactor, placed inside a programmable electrical furnace. mole of Mo per second). A number of repeated runs under the same
The TPD profiles of CO, CO2 and H2 O were obtained from room conditions were performed to ensure satisfactory reproducibility of
temperature to 1040 ◦ C, at 10 ◦ C min−1 under helium (AGA Chile, the data. The uncertainty in the calculation of reaction rates from
99.995%) flow of 50 mL min−1 . Evolution of desorbed gases was the GC peaks is 3%. The phenol/catechol ratios were determined at
monitored by a thermal conductivity detector (TCD). In addition, 10% conversion of guaiacol.
to quantify the gases produced during thermal decomposition of The stability of selected nitride catalysts during the HDO reac-
the surface functionalities, TPD was coupled with mass spectrom- tion was compared to that of a commercial reference catalyst
etry (Altamira AMI-200 R-HP instrument equipped with a SRS in a stainless steel continuous-flow micro-reactor. In a typi-
RGA-300 mass spectrometer). About 0.2 g of the activated carbon cal experiment, approximately 0.2 g of catalyst was diluted 1:1
114 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Fig. 1. Pore size distribution of HNO3 treated activated carbon supports.

with SiC (Soviquim, Chile) and loaded into the reactor tube, The BET surface areas, total and micropore volumes of supports
while the remaining reactor space was packed with SiC. Prior to and nitride catalysts are presented in Table 1. Significant differ-
the reaction, the nitride catalyst was reduced in situ using the ences in these textural parameters among the four supports are
same pretreatment conditions employed for the batch reaction, obvious. The CGran and GCA carbons both have higher surface areas
while the commercial Ni–Mo/Al2 O3 catalyst (Procatalyse, HR 346, but the latter has a larger fraction of micropores (76 vs. 32%). The
SBET = 256 m2 g−1 ) was sulfided in situ using a 10 vol.% H2 S (AGA lowest SBET displayed by the Darco support is consistent with this
Chile, 99.99%) in H2 , at a flow rate of 67.5 mL min−1 and 350 ◦ C for material having the lowest microporosity (19%) of all the supports.
3 h. The liquid reactant mixture and hydrogen were connected to The porosity of the GAC support is more evenly distributed: 57%
the reactor inlet where they flowed downward through the cata- mesoporosity and 43% microporosity.
lyst bed. The conditions for HDO reactions were as follows: 300 ◦ C, Oxidation pretreatment of the supports with HNO3 produced
3 MPa, 5.4 g/h of liquid feed corresponding to liquid hourly space varying changes in the textural properties of the original samples.
velocity (LHSV) of 27 h−1 , H2 gas hourly space velocity (GHSV) In particular, a significant loss of surface area (28%), total pore vol-
of 3600 h−1 , H2 /guaiacol molar ratio of 23. Fresh samples were ume (25%) and micropore volume (24%) was found for the CGran
collected at an hourly interval for 8–9 h with regular flushing pre- activated carbon; this is attributed to an increase in the quantity of
ceding each collection. The liquid products were then analyzed by oxygen-containing surface groups on the pore walls and entrances,
GC–FID. making them inaccessible to N2 molecules at 77 K [39]. The pre-
treatment resulted in significantly less micropore volume decrease
3. Results and discussion in the GAC carbon (5%), corresponding to a 4% loss of surface area; in
contrast, there was an increase in surface area and pore volume for
3.1. Textural properties the Darco and GCA activated carbons, possibly due to the removal
of impurities from the pores. The pore size distributions of the car-
The pore size distributions (PSD) for the activated carbons, cal- bons were not significantly modified after the treatment, however.
culated from the adsorption branch of the isotherm using the NLDFT Results summarized in Table 1 also show that impregnation of the
method, are shown in Fig. 1. All the support materials have a wide supports with Mo and Mo/Co, followed by thermal conversion to
range of pore sizes, including micropores in the range 0.4–2.0 nm. the nitrides, led to a general decrease in surface area, as well as total
However, the results for CGran and Darco supports reveal a pre- and micropore volume, especially in the case of the CGran support.
dominance of larger mesopores (up to 100 nm). In contrast, the GCA These results suggest that nitride formation occurred inside the
and GAC supports are more microporous; the latter also possesses porous structure. For the Darco support in particular, they suggest
an appreciable amount of larger mesopores (3–100 nm). preferential nitride formation inside the mesopores; conversely, for
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 115

Table 1
Nitrogen porosimetry results for supports and catalysts.

Sample SBET (m2 g−1 ) dpore (nm) Pore volume (cm3 g−1 )

TPV V

Darcoas-received 612 2.0 0.62 0.12


Darcopretreated 664 2.0 0.68 0.14
MoN/Darco-A 561 2.1 0.58 0.13
MoN/Darco-NH 560 2.1 0.60 0.14
CoMoN/Darco-A 475 2.3 0.54 0.10
CGranas-received 1402 1.6 1.15 0.37
CGranpretreated 1014 1.7 0.86 0.28
MoN/CGran-A 566 1.6 0.46 0.15
MoN/CGran-NH 571 1.6 0.47 0.15
CoMoN/CGran-A 461 1.9 0.44 0.11
GACas-received 976 1.1 0.56 0.32
GACpretreated 942 1.2 0.55 0.31
MoN/GAC-A 775 1.2 0.46 0.25
MoN/GAC-NH 752 1.2 0.45 0.24
CoMoN/GAC-A 706 1.5 0.52 0.21
GCAas-received 1132 0.9 0.51 0.39
GCApretreated 1202 0.9 0.55 0.42
MoN/GCA-A 995 0.9 0.45 0.35
MoN/GCA-NH 1066 0.9 0.49 0.35
CoMoN/GCA-A 950 0.9 0.44 0.33

the CGran support they suggest preferential deposition in microp- contained the greatest amount of CO2 - and CO-desorbing groups
ores or at the entrances to such pores. among the activated carbon supports. Table 2 also shows that the
GCA and Darco carbons had similar quantities of CO2 - and CO-
3.2. Surface chemical properties of the support releasing functional groups. (The H2 O profiles at low temperature,
indicating weakly bound desorbed molecules, were not quantified.)
The chemical nature of oxygen functionalities present on the Furthermore, comparison of TPD/MS results of pretreated and as-
activated carbon surface after HNO3 treatment was determined received activated carbon supports (not shown) indicates greater
from TPD/MS measurements, as shown in Fig. 2 and summarized amounts of CO2 - and CO-desorbing groups in the former. This is
in Table 2: lactonic (190–650 ◦ C) [40–42], carboxylic (200–300 ◦ C) consistent with the well documented effectiveness of oxidative
[41,42], phenolic (600–700 ◦ C) [40,42], carbonyl (800–980 ◦ C) HNO3 treatment [21]. It should be clarified, however, that if advan-
[40,42], and quinone groups (700–1000 ◦ C) [42,43]. Decomposi- tage is not taken of the surface oxygen anchoring sites – by making
tion of groups whose carbon atom is bonded to two oxygen atoms sure, during catalyst preparation, that there is attraction between
(carboxylic acids, lactones, and carboxylic anhydrides) releases the positively charged support surface and the negatively charged
CO2 , indicative of the presence of strong acidic sites [21,24]. All catalyst precursor, or vice versa – then the only benefit of surface
the supports exhibited pronounced peaks at low temperatures oxidation can be to render the support more hydrophilic (unless
(250–400 ◦ C) and thus possessed large amount of acidic CO2 - textural properties are significantly modified).
desorbing groups, as expected [24]. In the high temperature region, Fig. 3 shows the FT-IR spectra of the activated carbon sup-
the GCA, GAC and Darco supports presented broader 515–1000 ◦ C ports. The supports displayed bands in the 3500–3300 cm−1 ,
shoulders, which is indicative of the prevalence of phenolic and 1600–1500 cm−1 and 1500–1000 cm−1 regions which can be
quinone groups. The CGran support exhibited a 415–730 ◦ C peak assigned to anhydrides, quinonic, carboxylic or ether groups,
which is suggestive of the presence of mostly phenolic groups; respectively [44]. There is no apparent difference between the
decomposition of these groups, which leads to desorption of CO, GCA, GAC and Darco spectra, indicating, in agreement with TPD
indicates the presence of weakly acidic, neutral and basic groups results, that they have similar surface oxygenated species. For the
whose carbon atom is bonded to one oxygen atom [21,24]. In addi-
tion, integration of the evolved gas TPD profiles shows that CGran

Fig. 2. TPD profiles of the activated carbon supports. Fig. 3. FT-IR spectra of the activated carbon supports.
116 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Table 2
Surface chemical and acidic properties of oxidized supports.

Support TPD (arbitrary units) Acidity measurements

CO2 CO Acid strength (mV) Total acidity (mequiv. m−2 )

Darco 12 8 127 2.3


CGran 26 29 290 1.5
GAC 16 16 119 1.6
GCA 12 8 61 1.2

CGran support, a strong band appears at 1691 cm−1 which can (i.e. 0.5) and ␤-Mo2 N0.78 (i.e. 0.39) which suggests nitrogen enrich-
be assigned to lactonic groups [44], in agreement with the cor- ment, possibly residing in interstitial sites. Table 3 also shows that
responding TPD profile; another two bands were detected in the the N/Mo ratio was higher for the method 1 samples than method 2
3124–3037 cm−1 and 2904–2844 cm−1 regions, assigned to aro- samples; this suggests a higher degree of nitridation of the former.
matic and aliphatic groups [44]. By comparing the spectra in Fig. 3, The surface composition of the reduced and passivated nitride
CGran support displayed the greatest intensity of carboxylic or catalysts was determined by XPS and the results are summarized in
ether groups suggesting that it contains the greatest quantity of Table 4. The C 1s binding energies consisted of four peaks between
surface acidic oxygen groups. This result is consistent with the 284.8 and 289.2 eV. The peak at 284.8 eV was assigned to C C
interpretation from TPD profiles. and/or C C bonds of aromatic (and aliphatic) carbon [42,46], while
The surface acidity of the supports was estimated from poten- at 286.3 eV it is indicative of C O bonds in phenolic or ether groups
tiometric titration curves with n-butylamine as the probe molecule. [47,48], or may be due to the presence of C N bonds [49]. The peak
The results include the maximum acid strength of the surface at 287.7 eV is consistent with quinone-type groups or C N species
sites (derived from the initial electrode potential, E0 ) and the total [49,50], and at 289.3 eV with carboxyl groups and esters [48]. The
number of acid sites normalized by the surface area (acid site relative peak intensities (shown in parentheses) indicate the pre-
density). The acid strength can be determined according to the cri- dominance of aromatic carbon on the surface of all the catalysts.
terion proposed by Cid and Pecchi [38]: E0 > 100 mV, very strong The XRD patterns of the GAC-, GCA- and Darco-supported metal
sites; 0 < E0 < 100 mV, strong sites; −100 < E0 < 0 mV, weak sites; nitrides (see Supplementary data) revealed only peaks due to the
E0 < −100 mV, very weak sites. Accordingly, the results confirm that original carbon supports. The absence of Mo nitride diffraction
the HNO3 -treatment produced strong acid sites on the activated peaks suggests that the catalysts contained small crystallites of
carbon supports. Table 2 shows that the CGran, Darco and GAC sup- Mo nitride below the detection limit, or that the broad 0 0 2 and
ports displayed very strong acid sites with E0 > 100 mV, whereas 1 0 bands from the support may have masked the low-intensity
the GCA carbon displayed strong acid sites with 0 < E0 < 100 mV. Mo nitride peaks. The XRD results of CGran supported nitrides are
As mentioned, the Darco carbon had the highest acid site density, summarized in Fig. 4. The MoN/CGran-NH catalyst displayed the
while the GAC and CGran carbons had similar densities. The lowest characteristic peaks for ␤-Mo2 N0.78 (2 = 37.61◦ , 62.53◦ , 75.53◦ )
density of acid sites was measured for the GCA carbon. together with broad features (2 = 26◦ and 43◦ ) associated with the
carbon supports; a barely discernible peak at 2 = 37.41, consistent
3.3. Bulk and surface composition of nitrided catalysts with the ␥-Mo2 N (1 1 1) phase, was observed in the MoN/CGran-
A catalyst. Also shown in Fig. 4 is the diffraction pattern for
Bulk molybdenum, cobalt and nitrogen contents of passivated CoMoN/CGran-A, which indicated the presence of Co3 Mo3 N crys-
supported catalysts are listed in Table 3. The nitrogen content is tallites (2 = 40.09◦ , 42.59◦ , 46.59◦ ).
due to nitride formation and the creation of pyridine- and pyrrole- XPS results of Mo 3d5/2 , N 1s and Co 2p3/2 species are summa-
like functions during ammonolysis and N2 /H2 treatment [43]. To rized in Table 4. The binding energies of Mo 3d and N 1s of some
distinguish the nitrogen content due to the molybdenum nitride selected catalysts are shown in Fig. 5. The Mo 3d spectra, shown in
from those associated with the carbon support, we nitrided blank Fig. 5(a), presented two spectral lines centered at 232.6 ± 0.3 eV
support in a manner similar to that used to prepare the sup- and 235.7 ± 0.3 eV corresponding to the Mo 3d5/2 and Mo 3d3/2
ported Mo and CoMo nitride catalysts, and performed chemical spin orbits respectively, assigned to Mo(VI) species [51]. This indi-
nitrogen analysis: the results indicate that between 52 and 73% of cates that after the reduction treatment only Mo oxynitrides was
the total nitrogen content of the monometallic method 1 samples present on the catalyst surface. In the CoMo nitride catalysts the
were due to contributions from the support’s nitrogen-functional
groups; by comparison, the nitrogen species of the carbon support
accounted for between 27 and 45% of the total nitrogen content
of the monometallic method 2 catalyst; the corresponding sup-
port nitrogen groups for the bimetallic samples ranged 42–90%.
This clearly shows that for the same support the nitrogen contents
were consistently higher for samples prepared via method 1 (A-
series). Ammonolysis was thus more effective than with N2 /H2 . The
measured metal content values, in most part, correspond to theo-
retical values (10 wt% for Mo and 2.4 wt% for Co). However, some
of the catalysts had metal contents lower than the nominal values
which could be attributed to the possible loss of volatile molyb-
dates and cobaltates during the decomposition steps of catalyst
preparation such as drying and nitridation, a phenomenon which
has also been observed by other authors [22,45]. The bulk N/Mo
atomic ratios (due to the nitrogen structure of the molybdenum
nitride and denoted (N/Mo)nitride in Table 3) for the catalysts were
generally higher than the stoichiometric N/Mo ratio for ␥-Mo2 N Fig. 4. XRD patterns of CGran carbon-supported nitrides.
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 117

Table 3
Chemical analysis of metal nitride catalysts.

Catalyst Elemental composition (wt%) Atomic ratio

Mo Co N (N/Mo)total (N/Mo)nitride

MoN/Darco-A 10.20 – 3.70 2.5 0.8


MoN/Darco-NH 9.57 – 1.59 1.1 0.7
CoMoN/Darco-A 8.53 2.05 4.10 3.3 1.2
MoN/CGran-A 9.62 - 5.58 4.0 1.1
MoN/CGran-NH 10.3 – 1.26 0.8 0.5
CoMoN/CGran-A 10.1 2.87 4.55 3.1 0.3
MoN/GAC-A 7.6 – 2.14 1.9 0.9
MoN/GAC-NH 9.97 – 1.58 1.1 0.8
CoMoN/GAC-A 9.48 1.97 2.35 1.7 0.9
MoN/GCA-A 9.93 – 2.09 1.4 0.7
MoN/GCA-NH 8.79 – 1.16 0.9 0.6
CoMoN/GCA-A 8.44 2.03 2.54 2.1 1.2

Fig. 5. The XPS spectra of the (a) Mo 3d doublet and (b) N 1s in selected catalysts.
118 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Table 4
XPS results of reduced-passivated nitride catalysts.

Catalyst Binding energy (eV) (distribution/%) Surface atomic ratio (at.%)

C 1s Mo 3d5/2 N 1s Co 2p3/2 Mo/C Co/C (N/C)T (N/C)nitride a N/Mob

284.8 (72)
286.2 (16) 396.5 (19)
MoN/Darco-A 232.7 – 0.0140 – 0.0506 0.0096 0.7
287.7 (6) 398.4 (55)
289.3 (6) 400.1 (26)

284.8 (73)
286.2 (16) 396.9 (21)
MoN/Darco-NH 232.7 – 0.0159 – 0.0400 0.0084 0.5
287.7 (6) 398.5 (59)
289.3 (5) 400.1 (20)

284.8 (72)
286.2 (16) 396.5 (19)
CoMoN/Darco-A 232.6 781.4 0.0102 0.0061 0.0468 0.0089 0.9
287.7 (6) 398.4 (54)
289.3 (6) 400.1 (27)

284.8 (77)
286.3 (14) 396.8 (7)
MoN/CGran-A 232.8 – 0.0088 – 0.0531 0.0037 0.4
287.7 (5) 398.4 (63)
289.3 (4) 400.1 (30)

284.8 (79)
286.3 (13) 396.9 (12)
MoN/CGran-NH 233.0 – 0.0176 – 0.0529 0.0064 0.4
287.7 (4) 398.6 (55)
289.3 (4) 400.1 (33)

284.8 (80)
286.3 (12) 396.6 (17)
CoMoN/CGran-A 232.7 781.5 0.0182 0.0095 0.0088 0.0015 0.8
287.7 (4) 398.5 (57)
289.2 (4) 400.2 (26)

284.8 (76)
286.3 (16) 396.3 (15)
MoN/GAC-A 232.5 – 0.0128 – 0.0380 0.0057 0.5
287.7 (5) 398.3 (56)
289.2 (3) 399.6 (29)

284.8 (76)
286.3 (15) 396.8 (19)
MoN/GAC-NH 232.5 – 0.0157 – 0.0410 0.0078 0.5
287.7 (5) 398.4 (62)
289.2 (4) 400.1 (19)

284.8 (73)
286.2 (16) 396.3 (38)
CoMoN/GAC-A 232.5 781.3 0.0143 0.0075 0.0485 0.0184 1.3
287.7 (6) 398.3 (40)
289.3 (5) 399.6 (22)

284.8 (77)
286.2 (15) 396.3 (15)
MoN/GCA-A 232.4 – 0.0139 – 0.0436 0.0065 0.5
287.7 (5) 398.3 (56)
289.2 (3) 399.6 (29)

284.8 (76)
286.3 (15) 396.8 (19)
MoN/GCA-NH 232.5 – 0.0146 – 0.0379 0.0072 0.5
287.7 (5) 398.4 (62)
289.2 (4) 400.1 (19)

284.8 (72)
286.2 (17) 396.3 (38)
CoMoN/GCA-A 232.5 781.5 0.0104 0.0045 0.0421 0.0160 1.5
287.7 (6) 398.3 (40)
289.2 (5) 399.6 (22)
a
Calculated by multiplying the distribution of N 1s in the region of 396.3–396.9 eV to the total N/C atomic ratios.
b
N/Mo = (N/C)nitride /(Mo/C).

Co 2p3/2 binding energy was 781.4 eV; BE characteristic of Co(II) CGran support, suggesting that nitride forms inside the pores of
species is normally observed at 781.8 eV [52], showing a decrease the support rather than on the external surface of the supported
in positive Co2+ charge, indicating partial replacement of oxygen catalyst particles. This could be related to the heterogeneity in the
with a less electronegative nitrogen forming oxynitrides [53]. The distribution of oxygenated functionalities of the carbon support
result confirms that despite reduction of the passivated catalyst at during HNO3 -treatment. Due to diffusional effects the amount of
450 ◦ C, its surface is mainly an oxynitride rather than a nitride. The oxygen groups inside the pores is expected to be less than those
absence of nitride species on the external surface is attributed to a at the exterior [25]; consequently, the inner pores of the sup-
higher concentration of oxygenated functionalities on the support port are less positively charged and thus adsorbed less strongly to
surface (as a result of the HNO3 -pretreatment): negatively charged the heptamolybdate anion, leading to relatively easier decomposi-
heptamolybdate units interact more strongly with the positively tion and transformation of the precursor to nitride. This proposal
charged carbon surfaces which may have inhibited the formation goes further by demonstrating that the decrease in support textu-
of fully nitrided groups during nitridation. In contrast to these ral parameters after impregnation and thermal conversion shows
results, XRD analyses showed detectable nitride amounts on the that the XRD-detected Mo2 N, Mo2 N0.78 and Co3 Mo3 N species were
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 119

present on internal surfaces of the particles. Thus, the results imply


difference in composition between internal and external surfaces
of the carbon support.
The XPS peaks in the N 1s region shown in Fig. 5(b) and pre-
sented in Table 4, 398.3 ± 0.1, 399.6 and 400.1 eV, are ascribed to
pyridine, amide and pyrrolic groups created at the edges of the
graphene layers by nitridation [47,54]. The component at ∼396.5 eV
is attributed to N 1s in the Mo N bond [55], confirming the presence
of molybdenum oxynitrides. The relative abundance of each nitro-
gen species of the catalysts indicates that monometallic catalysts
prepared via method 2 contained more N species associated with
the Mo N bonds than catalysts prepared via method 1; the result
suggests that the effectiveness of surface molybdenum oxynitride
formation was related to the nitridation method. At the same time
higher total abundance of nitrogen-functional groups of the carbon
support was obtained for the method 1 samples, compared to the
method 2 samples. This result qualitatively agrees with the bulk
nitrogen content analysis in Table 3, confirming that ammonia is
more effective at nitriding the carbon support.
The atomic Mo/C, N/C and Co/C ratios are also listed in Table 4.
The XPS atomic ratios, estimated from intensity ratios of particle
related peaks and support related peaks, have a strong dependence
Fig. 6. Surface N/Mo atomic ratio (from XPS) vs. bulk N/Mo atomic ratio (from
on catalyst dispersion, and can qualitatively be used to compare
chemical analysis) of the carbon-supported catalysts.
the dispersion of supported catalysts [56]. The differences in Mo/C
ratios are indicative of dispersion differences when the catalysts are
prepared under different conditions. Thus, for example, the cata- A plot of the surface N/Mo atomic ratio (calculated from XPS) vs.
lysts prepared via method 2 (NH series) displayed higher Mo/C ratio bulk N/Mo atomic ratio (calculated from chemical analysis) after
than those prepared via method 1 (A-series). It can be concluded subtracting nitrogen content on the carbon support is shown in
from Table 4 that the thermal conversion of the oxide catalyst Fig. 6, and clearly indicates that for monometallic catalysts the bulk
precursor using N2 /H2 mixture led to more highly dispersed Mo remained enriched with nitrogen relative to the surface compo-
oxynitride particles. We can make a hypothesis on the basis of the sition. This result suggests that molybdenum nitrides/oxynitrides
information from the bulk and surface concentration of nitrogen with higher nitrogen deficiencies were located in the interior of the
moieties on the carbon support. Nitridation reduces the quantity of catalysts, in line with the interpretation derived from XRD and XPS.
surface oxygen groups and simultaneously increases the nitrogen In contrast, the CoMo nitride catalysts, except for CoMoN/CGran-A
functional groups; it has been confirmed that ammonolysis is more catalyst, had higher nitrogen concentration on the external surface
effective than N2 /H2 mixture for the nitridation of carbon supports. than in the bulk; this result is not yet clear.
Consequently, the greater quantity of negatively charged surface The textural and chemical surface properties of the support
nitrogen species in the method 1 samples would induce a greater could also influence Mo dispersion as demonstrated abundantly in
degree of partial distribution of molybdenum oxynitrides towards the literature [24]. Indeed, the highest dispersion displayed by the
the interior of the catalyst. The N/C atomic ratio, which expresses MoN/CGran-NH catalyst is attributed to the fact that this support
the dispersion of the total nitrogen species (including nitrogen from has the most abundant oxygen surface groups (most hydrophilic
oxynitride and nitrogen from organic species) on the carbon sup- character) and the highest mesoporosity; this facilitates access of
port surface, did not follow any observable trend; these results are the aqueous solution to the internal pore structure and allows a
less informative in relation to the degree of dispersion of the molyb- homogenous radial distribution of the metal precursor within the
denum oxynitride phase due to contributions from surface nitrogen particles of the support [23]. Conversely, the lowest dispersion of
species on the carbon support. However, the N/C atomic ratio per- the MoN/GCA-NH catalyst is attributed to the lowest concentra-
taining to the molybdenum oxynitride species can be elucidated tion of oxygen surface groups (most hydrophobic character) as well
by multiplying the total N/C atomic ratio by the percent distribu- as the predominance of micropores in the support. Furthermore,
tion, denoted as (N/C)nitride and shown in Table 4. Except for the the intermediate dispersion of the MoN/GAC-NH catalyst corre-
samples supported on Darco carbon, (N/C)nitride atomic ratios for lates with this solid possessing intermediate surface chemical and
monometallic samples are higher for method 2 samples than for textural property. It is important to remark that the presence of
method 1 samples. This result indicates the nitridation method is oxygen functional groups can render the support surface negatively
an essential factor in the control of the relative dispersion of molyb- charged over a wide range of pH condition, causing electrostatic
denum oxynitride species for carbon supported nitrides. The Co/C repulsions between the support and the heptamolybdate anion
ratio did not follow any observable trend. [25]. This can result in metal particles aggregation, leading to less
The surface N/Mo atomic ratio was evaluated from the XPS Mo/C surface Mo atoms as evident by the low Mo/C atomic ratio of the
and (N/C)nitride atomic ratios and is summarized in Table 4. The MoN/CGran-A catalyst.
calculated values were close to the stoichiometric N/Mo value of ␥-
Mo2 N and ␤-Mo2 N0.78 . However, because XPS results indicated the 3.4. Catalytic activity
presence of only molybdenum oxynitride, this observation is prob-
ably due to excess nitrogen residing in defect sites like irregular The conversion of guaiacol and the evolution of reaction prod-
grain boundary surfaces. Moreover, the N/Mo atomic ratio was sim- ucts are illustrated in Figs. 7 and 8. Periodic samplings of the
ilar to each other suggesting that the method of nitridation and the liquid mixture in the reactor were analyzed by GC, from which the
surface chemistry of the support had no observable control on the concentration of the reactant and the product yields were deter-
surface Mo oxynitride species formed. This infers that the nature mined relative to the hexadecane standard. The observed products
of the nitrogen deficient sites was the same for all the catalysts. were similar for all the catalysts and include phenol, catechol,
120 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Fig. 7. Variation of the transformation of guaiacol and the yield of products with time with (a) MoN/GCA-A, (b) MoN/GCA-NH, (c) CoMoN/GCA-A, (d) MoN/GAC-A (e)
MoN/GAC-NH, and (f) CoMoN/GAC-A catalysts.

deoxygenated compounds (such as cyclohexene, cyclohexane and Fig. 7(c) and (d), respectively. By comparison, the CGran-supported
benzene), and heavy compounds (such as mono- to tetra-methyl catalysts exhibited some slight differences in the products distri-
phenols and dimethyl catechols). Other possible products, such as bution at higher conversion, as shown in Fig. 8(a)–(c): phenol was
anisole, m- and o-cresol were not observed. Furthermore, CGran- found to be the dominant product over the MoN/CGran-NH cata-
and GAC-supported catalysts produced very small quantities of lyst while heavy products and phenol were the main competing
methylcatechol which were attributed to the higher total acidi- products formed in almost equal amounts with the MoN/CGran-A
ties of these catalysts. Fig. 7 illustrates similarities in the product and CoMoN/CGran-A catalysts. As seen in Fig. 8(d)–(f), the Darco-
distribution for the GCA-supported Mo nitride catalysts (Fig. 7(a) supported catalysts showed similar behavior in the evolution of
and (b)) and the GAC-supported Mo nitride catalysts (Fig. 7(d) products. The product evolution observed indicates that guaiacol
and (e)): guaiacol was mainly converted to phenol, while cate- HDO followed the reaction scheme proposed by Bui et al. [57]
chol and the deoxygenated compounds were observed in smaller shown in Fig. 9. The two general pathways are: (i) initial demethy-
quantities. Heavy compounds were formed in significant amount lation (DME) to form catechol, followed by dehydroxylation to form
and are attributed to the methylation of the aromatic ring [57]. phenol; or (ii) direct demethoxylation (DMO) to form phenol. (Light
The production of heavy compounds was even more prominent in products such as methane and methanol could not be separated
the CoMo nitride catalysts supported on GCA and GCA, as seen in by the column used under batch conditions, but they are expected
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 121

Fig. 8. Variation of the transformation of guaiacol and the yield of products with time with (a) MoN/CGran-A, (b) MoN/CGran-NH, (c) CoMoN/CGran-A, (d) MoN/Darco-A, (e)
MoN/Darco-NH, and (f) CoMoN/Darco-A catalysts.

byproducts of DME and DMO, respectively.) Further deoxygenation Table 5


Catalytic activities of carbon-supported Mo nitride catalysts.
of phenol produces benzene, cyclohexene and cyclohexane, while
methyl-substitution of catechol forms methylcatechol. Continuous Catalyst Activity Intrinsic activity
production of catechol at longer reaction times indicated that the (×106 mol g−1
catalyst
s−1 ) (×104 molec. Mo at−1 s−1 )
conversion of catechol to phenol was not prominent although both MoN/Darco-A 6.4 60.5
demethylation and direct demethoxylation occurred over these MoN/Darco-NH 6.5 65.1
catalysts. The same tendency was observed for bulk metal nitrides CoMoN/Darco-A 6.2 69.6
MoN/CGran-A 4.6 46.3
[30].
MoN/CGran-NH 9.0 83.6
Table 5 summarizes the catalytic activities. The initial reaction CoMoN/CGran-A 5.3 49.9
rates calculated from the slopes of the guaiacol conversion curves MoN/GAC-A 5.9 74.1
are given in Fig. 10. Negligible conversions were obtained using MoN/GAC-NH 7.9 75.6
the bare supports (not shown), indicating that the activities were CoMoN/GAC-A 2.9 33.1
MoN/GCA-A 6.8 65.8
associated with the Mo oxynitride and not the support. Thus, the
MoN/GCA-NH 7.5 81.8
surface oxygen functional groups of the activated carbon did not CoMoN/GCA-A 5.4 61.2
participate in the conversion of guaiacol; however, the interaction
122 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Fig. 9. Hydrodeoxygenation pathway of guaiacol.


Adapted from Bui et al. [57].

between the metal precursor compound and the surface groups of Ni–W/carbon catalysts [22,23]. The higher HDO activity for method
the support promotes good dispersion, leading to enhanced activ- 2 catalysts coupled with the reported advantages in the large-scale
ities. This is supported by the observed higher activities compared synthesis of Mo nitride using a N2 /H2 mixture as a reactant over the
to our previous results for non-modified carbon supported Mo2 N NH3 synthesis makes method 2 attractive for potential industrial
catalysts [27]. The reaction rate was not related to the surface N/Mo application [34]; thus, for example, the N2 and H2 reactants can be
atomic ratio, as a consequence of the catalysts possessing the same economically recycled by drying, and their use simplifies handling
nature of surface nitrogen deficient sites discussed previously in procedures as well as eliminate heat transfer problems associated
regard to the reported XPS N/Mo ratio data. However, the reported with endothermic decomposition of NH3 [34].
activities were affected by the method of nitridation: the cata- Other trends can be observed in Fig. 10 when the catalysts
lysts prepared by method 2 (NH series) had consistently higher prepared using the same method but dispersed on different sup-
activities. This is related to their higher Mo oxynitride dispersion ports are considered. The specific activity of the method 2 catalysts
(illustrated in Fig. 11), confirmed by the Mo/C and (N/C)nitride atomic (NH series) decreased in the order: MoN/CGran-NH > MoN/GAC-
ratio in Table 4. The method 2 catalysts contained a greater number NH > MoN/GCA-NH > MoN/Darco-NH. Thus HDO activity appears
of active sites associated with molybdenum oxynitride, leading to to be favored by a combination of higher Mo dispersion and
higher activities. Correlation of HDO activity with dispersion has higher support mesoporosity (MoN/Darco-NH being the excep-
similarly been reported for sulfided Co–Mo/carbon and reduced tion): the highest Mo oxynitride dispersion and the easiest
reactant accessibility to the support’s mesoporous structure led

Fig. 10. Reaction rates of carbon-supported Mo nitride catalysts. Fig. 11. Reaction rates vs. XPS Mo/C atomic ratio.
I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124 123

to the highest HDO activity of the MoN/CGran-NH catalyst.


The reaction rates of the method 1 catalysts (A-series) corre-
late with Mo oxynitride dispersion and decrease in the order:
MoN/GCA-A > MoN/GAC-A > MoN/Darco > MoN/CGran-A. Some of
the Darco-supported catalysts exhibited inferior activity, compared
to GCA-supported catalysts, despite their higher dispersion and
higher mesoporosity. This surprising behavior could be due to an
overestimation of the Mo signal obtained by XPS. Considering that
the Darco support possessed the lowest surface area of all the sup-
ports, and that all the catalysts were impregnated with a similar Mo
content, this catalyst should contain the largest Mo nitride particle
sizes. However, their measured Mo/C atomic surface ratios were
higher probably due to the inability of X-ray photons to penetrate
larger Mo nitride particles, leading to a higher intensity of the Mo
3d XPS signal and thus an overestimation of the Mo/C atomic sur- Fig. 12. Phenol/catechol ratio calculated at 10% guaiacol conversion.
face ratios. Similar behavior was previously observed by Lagos et al.
[58]. Therefore, the low activity of Mo nitride catalysts supported
nature of reduced passivated supported catalysts is inevitably
on Darco carbon was probably due to the loss of active sites through
different due to the elimination of the weakly adsorbed NHx species
the formation of agglomerates.
by passivation; thus, we infer that the surface of the carbon-
The intrinsic activities based on the Mo content are also given
supported Mo nitride catalysts expose coordinately unsaturated
in Table 5. The trends were similar to the total reaction rates.
Mo atoms and 4-fold type nitrogen deficient sites. Hypothesis can
Figs. 10 and 11 also show that the addition of Co did not increase the
be made that coordinately unsaturated Mo atoms are responsible
activity of the catalysts. In fact, GCA- and GAC-supported CoMoN
for demethoxylation (Caromatic OCH3 bond cleavage) and dehy-
catalysts were not nearly as active as their MoN counterparts. In
droxylation while demethylation (Cmethyl O bond cleavage) occurs
our previous, it was shown that addition of Co did not enhance the
at nitrogen deficient sites. This hypothesis is analogous to the
activity of unsupported Mo nitride catalyst [30]. This was attributed
distinction between sites for sulfide catalysts reported by Fer-
to incomplete formation of the bimetallic nitride, Co3 Mo3 N, phase.
rari et al. [60]. All the catalysts displayed high phenol production
The same interpretation applies to supported catalysts as well.
and this cannot be related directly to the surface acidic proper-
Unlike bimetallic sulfides and oxides, the formation of single-phase
ties of the carbon supports; recently, Sepulveda et al. [61] showed
bimetallic nitride is not trivial and often exists in multiple phases.
that strong acid sites favor catechol formation. The preference
Synthesis of pure-phase bimetallic nitride catalysts (Co3 Mo3 N and
for the demethoxylation pathway (Caromatic OCH3 bond cleavage)
Ni2 Mo3 N) and its HDO catalysis are warranted. Moreover, the sur-
suggests that unsaturated Mo sites were predominant in the cata-
face Mo/Co atomic ratio (from XPS analysis) is higher than the bulk
lyst. However, these high phenol/catechol ratios were significantly
Mo/Co atomic ratio (from metal content analysis), indicating that
lower than those observed for unsupported nitride catalysts [30],
cobalt is preferentially deposited in the interior of the support and
suggesting that the active sites on nitrides and/or oxynitrides were
are possibly trapped underneath molybdenum. This can be taken to
modified by the support, rendering them less selective for the
suggest that during impregnation cobalt migrated into the interior
demethoxylation route compared to the unsupported catalysts.
of the catalyst and caused the partial migration of molybdenum
This confirms that the generation of oxygen groups on the carbon
towards the exterior, in good agreement with the results obtained
surface has an indirect contribution on phenol/catechol selectivity
by Ferrari et al. [17]. This inhomogeneity in the distribution of cobalt
through the precursor/support interaction, modifying the nature of
could explain the diminishing influence on activity after Co incor-
the active sites and their selectivity. For example, N atoms present
poration: a higher concentration of surface Co species could create
on the surface of bulk Mo nitride catalysts may not be there on sup-
new (or modify) active sites which could enhance activity; con-
ported catalysts due to the metal nitride–support interaction and
versely, a large proportion of Co in the pores will not only hinder
this can have an overall effect on the phenol/catechol ratio and the
their accessibility but may also limit reactant diffusion into other
amount of deoxygenated products. Fig. 12 does not reveal any clear
catalytic active sites located in the internal surfaces of the support
tendencies in the phenol/catechol ratio, suggesting that the extent
which will negatively affect reaction rates. However, a more exten-
and manner of the effect of the surface chemistry of the support,
sive analysis of promoter effect (utilizing different Co precursors
dictated by the surface oxygen groups, is unclear. Therefore, this
and preparative methods such as sequential vs. co-impregnation)
warrants additional research.
on activity requires further investigation.
The selectivity results are summarized in Fig. 12. In con-
trast to metal hydrogenation catalysts like Ru [4], metal sulfide
and metal nitride hydrotreating catalysts have a higher selectiv-
ity for HDO reactions relative to hydrogenation of aromatic and
olefinic compounds [8,30]. During conversion of guaiacol over
Mo nitride/carbon catalysts, both demethylation and demethoxy-
lation reactions take place only at the active sites situated on
the metal nitrides due to the inertness of the carbon support.
Blank experiments with all the supports resulted in negligi-
ble conversion, suggesting that the physicochemical differences
of the carbon supports had no direct influence on the product
selectivity. The dual-pathway behavior of the Mo2 N/carbon cat-
alytic system provides evidence that two kinds of active sites
are present on these catalysts. Theoretical studies reveal that the
surface of fresh bulk Mo2 N consists of coordinately unsaturated Fig. 13. Time-on-stream behavior of the MoN/GAC-NH catalysts in terms of total
Mo and N atoms, and 4-fold type vacancies [59]. The surface conversion for HDO of guaiacol at 300 ◦ C, 3 MPa H2 pressure, H2 /guaiacol ratio of 23.
124 I.T. Ghampson et al. / Applied Catalysis A: General 439–440 (2012) 111–124

Developing a robust HDO catalyst for pyrolysis oil upgrading [10] Y.-C. Lin, C.-L. Li, H.-P. Wan, H.-T. Lee, C.-F. Liu, Energy Fuels 25 (2011) 890–896.
is a considerable challenge. Possible reasons for catalyst deacti- [11] E. Laurent, B. Delmon, Appl. Catal. A 109 (1994) 97–115.
[12] E. Laurent, B. Delmon, J. Catal. 146 (1994) 281–291.
vation during HDO include coking, poisoning and loss of active [13] A. Popov, E. Kondratieva, J.M. Goupil, L. Mariey, P. Bazin, J.-P. Gilson, A. Travert,
sites through surface chemistry changes [11,12]. A preliminary F. MaugeÌ, J. Phys. Chem. C 114 (2010) 15661–15670.
investigation of time-on-stream behavior of our nitrided catalysts, [14] A. Centeno, E. Laurent, B. Delmon, J. Catal. 154 (1995) 288–298.
[15] P.E. Ruiz, K. Leiva, R. Garcia, P. Reyes, J.L.G. Fierro, N. Escalona, Appl. Catal. A
compared to a reference commercial sulfided NiMo/Al2 O3 catalyst, 384 (2010) 78–83.
was conducted in a continuous flow reactor. It is summarized in [16] M. Ferrari, R. Maggi, B. Delmon, P. Grange, J. Catal. 198 (2001) 47–55.
Fig. 13 and agrees with the results of Monnier et al. [62]. The liquid [17] M. Ferrari, B. Delmon, P. Grange, Carbon 40 (2002) 497–511.
[18] V.N. Bui, D. Laurenti, P. Delichère, C. Geantet, Appl. Catal. B 101 (2011) 246–255.
flow rate was chosen to obtain low conversion. The nitrided cata-
[19] A. Centeno, V. CH, R. Maggi, B. Delmon, in: A.V. Bridgwater (Ed.), Developments
lysts displayed higher stability than the sulfided catalyst after 4 h in Thermochemical Biomass Conversion, Blackie Academic & Professional, Lon-
on stream under continuous operation. The gradual deactivation of don, 1997, pp. 602–610.
[20] A. Centeno, O. David, C. Vanbellinghen, R. Maggi, B. Delmon, in: A.V. Bridgwater
the sulfided NiMo/Al2 O3 catalyst could be due to loss of the sulfide
(Ed.), Developments in Thermochemical Biomass Conversion, Blackie Academic
phase during HDO reaction. & Professional, London, 1997, pp. 589–601.
[21] G. de la Puente, A. Centeno, A. Gil, P. Grange, J. Colloid Interface Sci. 202 (1998)
155–166.
4. Conclusions
[22] S. Echeandia, P.L. Arias, V.L. Barrio, B. Pawelec, J.L.G. Fierro, Appl. Catal. B 101
(2010) 1–12.
Use of activated carbon materials with different textural and [23] G. de la Puente, A. Gil, J.J. Pis, P. Grange, Langmuir 15 (1999) 5800–5806.
chemical surface properties to prepare supported Mo nitride cat- [24] L.R. Radovic, F. Rodriguez-Reinoso, in: P.A. Thrower (Ed.), Chem. Phys. Carbon,
Marcel Dekker, 1997, pp. 243–358.
alysts resulted in their different activities in HDO of guaiacol, [25] G. de la Puente, J.A. Menendez, Solid State Ionics 112 (1998) 103–111.
demonstrating rapid production of significant amounts of phe- [26] T. Nimmanwudipong, R. Runnebaum, D. Block, B. Gates, Catal. Lett. 141 (2011)
nol. This indicated that the transformation of guaiacol proceeded 779–783.
[27] C. Sepúlveda, K. Leiva, R. García, L.R. Radovic, I.T. Ghampson, W.J. DeSisto, J.L.G.
mostly through the direct demethoxylation route, bypassing the Fierro, N. Escalona, Catal. Today 172 (2011) 232–239.
formation of catechol. The higher specific activity of carbon- [28] S.T. Oyama, Catal. Today 15 (1992) 179–200.
supported catalysts prepared by using a N2 /H2 mixture, compared [29] J.G. Chen, Chem. Rev. 96 (1996) 1477–1498.
[30] I.T. Ghampson, C. Sepulveda, R. Garcia, B.G. Frederick, M.C. Wheeler, N. Escalona,
to similarly supported Mo nitrides prepared by ammonolysis, was W.J. DeSisto, Appl. Catal. A 413–414 (2012) 78–84.
attributed to a higher dispersion of Mo oxynitride. The dispersion [31] Y. Li, Y. Zhang, R. Raval, C. Li, R. Zhai, Q. Xin, Catal. Lett. 48 (1997) 239–245.
was related to the surface chemistry and the textural properties [32] J.W. Logan, J.L. Heiser, K.R. McCrea, B.D. Gates, M.E. Bussell, Catal. Lett. 56 (1998)
165–171.
of the support: high concentration of oxygen surface groups on the
[33] L. Volpe, M. Boudart, J. Solid State Chem. 59 (1985) 332–347.
support and high mesoporosity of the support promotes better dis- [34] R.S. Wise, E.J. Markel, J. Catal. 145 (1994) 344–355.
persion. The most active HDO catalyst (MoN/CGran-NH) was the [35] J.P. Olivier, Carbon 36 (1998) 1469–1472.
[36] M. Kruk, M. Jaroniec, K.P. Gadkaree, J. Colloid Interface Sci. 192 (1997) 250–256.
one that contained highly exposed Mo species on a highly meso-
[37] C.D. Wagner, L.E. Davis, M.V. Zeller, J.A. Taylor, R.H. Raymond, L.H. Gale, Surf.
porous support. Surprisingly, a generally diminishing influence on Interface Anal. 3 (1981) 211–225.
activity was observed after incorporation of Co to prepare bimetal- [38] R. Cid, G. Pecchi, Appl. Catal. 14 (1985) 15–21.
lic nitrided catalysts. [39] H.P. Boehm, Carbon 32 (1994) 759–769.
[40] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. Órfão, Carbon 37 (1999)
1379–1389.
Acknowledgements [41] Y. Otake, R.G. Jenkins, Carbon 31 (1993) 109–121.
[42] U. Zielke, K.J. Hüttinger, W.P. Hoffman, Carbon 34 (1996) 983–998.
[43] J.M. Calo, D. Cazorla-Amorós, A. Linares-Solano, M.C. Román-Martínez, C.S.-M.
The authors acknowledge the financial support of DOE Epscor De Lecea, Carbon 35 (1997) 543–554.
Grant #DE-FG02-07ER46373 and the financial support from CON- [44] P.E. Fanning, M.A. Vannice, Carbon 31 (1993) 721–730.
ICYT Chile, projects PFB-27, PIA-ACT-130 and FONDECYT No. [45] G.M. Dolce, P.E. Savage, L.T. Thompson, Energy Fuels 11 (1997) 668–675.
[46] J.P.R. Vissers, S.M.A.M. Bouwens, V.H.J. de Beer, R. Prins, Carbon 25 (1987)
1100512 grants. I. Tyrone Ghampson is indebted to NSF Career 485–493.
Award 0547103 for sponsoring a trip to the University of Concep- [47] S. Biniak, G. Szymanski, J. Siedlewski, A. Swiatkowski, Carbon 35 (1997)
ción. The authors also gratefully acknowledge valuable discussions 1799–1810.
[48] S.D. Gardner, C.S.K. Singamsetty, G.L. Booth, G.-R. He, C.U. Pittman, Carbon 33
on N2 sorption with Rachel Pollock, and the technical assistance of (1995) 587–595.
Nick Hill and Manuel Veliz. [49] E. Riedo, F. Comin, J. Chevrier, F. Schmithusen, S. Decossas, M. Sancrotti, Surf.
Coat. Technol. 125 (2000) 124–128.
[50] E. D’Anna, M.L. De Giorgi, A. Luches, M. Martino, A. Perrone, A. Zocco, Thin Solid
Appendix A. Supplementary data Films 347 (1999) 72–77.
[51] S.W. Yang, C. Li, J. Xu, Q. Xin, J. Phys. Chem. B 102 (1998) 6986–6993.
Supplementary data associated with this article can be [52] J.S. Girardon, E. Quinet, A. Griboval-Constant, P.A. Chernavskii, L. Gengembre,
A.Y. Khodakov, J. Catal. 248 (2007) 143–157.
found, in the online version, at http://dx.doi.org/10.1016/
[53] M.L. Kaliya, S.B. Kogan, Catal. Today 106 (2005) 95–98.
j.apcata.2012.06.047. [54] R.J.J. Jansen, H. van Bekkum, Carbon 33 (1995) 1021–1027.
[55] K. Hada, M. Nagai, S. Omi, J. Phys. Chem. B 105 (2001) 4084–4093.
[56] A.M. Venezia, Catal. Today 77 (2003) 359–370.
References
[57] V.N. Bui, G. Toussaint, D. Laurenti, C. Mirodatos, C. Geantet, Catal. Today 143
(2009) 172–178.
[1] A.V. Bridgwater, D. Meier, D. Radlein, Org. Geochem. 30 (1999) 1479–1493. [58] G. Lagos, R. García, A.L. Agudo, M. Yates, J.L.G. Fierro, F.J. Gil-Llambías, N.
[2] D.C. Elliott, Energy Fuels 21 (2007) 1792–1815. Escalona, Appl. Catal. A 358 (2009) 26–31.
[3] E. Furimsky, Appl. Catal. A 199 (2000) 147–190. [59] G. Frapper, M. Pelissier, J. Hafner, J. Phys. Chem. B 104 (2000) 11972–11976.
[4] D.C. Elliott, T.R. Hart, Energy Fuels 23 (2008) 631–637. [60] M. Ferrari, S. Bosmans, R. Maggi, B. Delmon, P. Grange, Catal. Today 65 (2001)
[5] E. Laurent, B. Delmon, Appl. Catal. A 109 (1994) 77–96. 257–264.
[6] J. Zakzeski, P.C.A. Bruijnincx, A.L. Jongerius, B.M. Weckhuysen, Chem. Rev. 110 [61] C. Sepúlveda, N. Escalona, R. Garcia, D. Laurenti, M. Vrinat, First International
(2010) 3552–3599. Congress on Catalysis for Biorefineries (CatBior), Torremolinos-Málaga, Spain,
[7] H.Y. Zhao, D. Li, P. Bui, S.T. Oyama, Appl. Catal. A 391 (2011) 305–310. 2011.
[8] V.N. Bui, D. Laurenti, P. Afanasiev, C. Geantet, Appl. Catal. B 101 (2011) 239–245. [62] J. Monnier, H. Sulimma, A. Dalai, G. Caravaggio, Appl. Catal. A 382 (2010)
[9] A. Gutierrez, R.K. Kaila, M.L. Honkela, R. Slioor, A.O.I. Krause, Catal. Today 147 176–180.
(2009) 239–246.

You might also like