Você está na página 1de 8

Article

Cite This: Energy Fuels 2017, 31, 10858-10865 pubs.acs.org/EF

Catalytic Hydrodeoxygenation of Guaiacol over Palladium Catalyst


on Different Titania Supports
Mohong Lu,†,‡ Hu Du,† Bin Wei,† Jie Zhu,† Mingshi Li,*,† Yuhua Shan,† and Chunshan Song*,‡

Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology and Advanced Catalysis and Green Manufacturing
Collaborative Innovation Center, Changzhou University, Changzhou, Jiangsu 213164, PR China

Clean Fuels and Catalysis Program, EMS Energy Institute, Department of Energy & Mineral Engineering and Department of
Chemical Engineering, Pennsylvania State University, 209 Academic Projects Building, University Park, Pennsylvania 16802, United
States

ABSTRACT: Pd catalysts supported on TiO2 with different crystalline phases were prepared with formaldehyde as reducing
agent and examined for hydrodeoxygenation (HDO) of guaiacol. Their properties were characterized by N2 adsorption, X-ray
diffraction, transmission electron microscopy, and X-ray photoelectron spectroscopy. Compared to the carbon-supported Pd
catalysts, TiO2-supported Pd catalysts exhibited higher C−O bond scission ability, which may be attributed to the presence of
partially reduced titanium species originating from the reduction of Ti4+ by spillover hydrogen from Pd at 200 °C on the surface
of TiO2. Guaiacol was hydrogenated on Pd sites to give 2-methoxycyclohexanol, which diffused to partially reduced titanium
species and subsequently reacted with hydrogen from Pd to generate cyclohexane. Anatase TiO2-supported Pd catalyst gave the
highest HDO activity of guaiacol among the Pd catalysts supported on three types of TiO2 (anatase, rutile, and their mix, P25),
suggesting that more partially reduced titanium species are in favor of the HDO reaction because anatase is facile to reduce by H2
at 200 °C. Higher selectivity of cyclohexane for Pd/TiO2 reduced at 500 °C than that reduced at 200 °C further confirmed that
the enhanced C−O bond scission ability of Pd/TiO2 is mainly attributed to the partially reduced titanium species on the surface
of TiO2.

1. INTRODUCTION Recently noble metals, such as Pd, Rh, Pt, and Ru and their
The concerns for energy security and global environmental alloys, have received increasing attention as an effective
alternative to sulfided catalysts in HDO of bio-oils.26−31
issues have motivated increasing research efforts on biomass as
Supported noble metal catalysts can have high HDO activity
a promising renewable resource.1 The development of pyrolysis
at mild reaction conditions. Supports of noble metal catalysts
technology for transforming biomass into bio-oils has made the
have remarkable influence on the HDO performance. Acidic
use of biomass possible.2,3 Bio-oils made from biomass,
supports, such as γ-Al2O3, SiO2−Al2O3, and acid-treated carbon
however, cannot be utilized due to their high viscosity, low were considered as support for noble metal catalysts in the
heating value, and low thermal and chemical stabilities caused HDO of bio-oils. A bifunctional mechanism was proposed for
by high content of oxygen.4−7 Therefore, it is necessary for bio- HDO of guaiacol over the noble metal catalysts supported on
oil to be upgraded so that it can be used as a standard fuel. acidic oxide.26 The benzene ring of guaiacol was hydrogenated
Catalytic hydrodeoxygenation (HDO) is considered to be an on noble metal sites and then deoxygenation occurred on acidic
efficient technology to upgrade bio-oils, during which the sites of the support. The transalkylation activity of catalysts was
oxygen in bio-oils is removed by reaction with hydrogen gas to dramatically influenced by the type of acidic sites by
give water in the presence of catalysts without unnecessary loss comparison of catalytic HDO over Pt/γ-Al2O3 and Pt/HY.27
of carbon.8,9 Wang et al. studied the Pd−Fe supported on carbon catalyst for
Supported CoMoS and NiMoS catalysts which exhibit high HDO of guaiacol and demonstrated that the synergy between
activities in the hydrodesulfurization (HDS) and hydro- Pd and Fe contributed to the high selectivity to benzene.30
denitrogenation (HDN) process were examined for the HDO Palladium is commonly used as a heterogeneous catalyst for
of bio-oils in earlier research.10−14 The active sulfided state of selective hydrogenation.32−35 The catalytic performance of Pd
catalyst can be maintained by the sulfur from fossil fuel. It is metal is significantly affected by the interaction between Pd
difficult, however, for the sulfided catalysts to keep active in particles and the supports.32−35 TiO2 has been widely
HDO of bio-oils due to the low content of sulfur (typically < investigated as a support of noble metal catalyst due to its
0.1 wt %) in the bio-oil except when additional sulfur was reducible property. It is well-known that the strong metal
added into the feed.15,16 A number of catalysts were explored to support interaction (SMSI) between Pd and TiO2 occurred
replace the sulfided catalysts for HDO of bio-oils. Transition when Pd/TiO2 was reduced in hydrogen gas at high
metal phosphides,17−19 carbides,20−22 and nitrides23−25 have
been widely investigated because they exhibit excellent HDO Received: May 27, 2017
performance. However, fast deactivation of these catalysts Revised: July 14, 2017
during the HDO process is still an important problem. Published: August 31, 2017

© 2017 American Chemical Society 10858 DOI: 10.1021/acs.energyfuels.7b01498


Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

temperature. Partially reduced TiOx (x < 2) could migrate onto °C. The mixture was filtered, washed and dried in a vacuum oven at 60
the surface of Pd and cover Pd particles, which will change the °C for 12 h. The obtained Pd/TiO2 (anatase), Pd/TiO2 (rutile), Pd/
adsorption capacity and catalytic activity of Pd.36−38 Crossley’s TiO2 (P25), and Pd/mesoporous carbon and Pd/activated carbon are
group compared the guaiacol conversion on Ru metal referred to as Pd/TiO2(A), Pd/TiO2(R), Pd/TiO2(P), Pd/MC, and
Pd/AC, respectively.
supported on TiO2 and nonreducible C, SiO2, and Al2O3 and 2.2. Catalyst Characterization. The adsorption of nitrogen was
found that Ru/TiO 2 was more active than on other performed using a Micromeritics ASAP 2020 automated system at
supports.39,40 They suggested that the high activity of the −196 °C to obtain the BET specific surface area, pore volume, and
Ru/TiO2 catalyst is attributed to spillover of hydrogen from the average pore diameter. The crystal structure of the catalysts were
Ru metal to the reducible TiO2 to produce defect sites. The conducted by X-ray diffraction (XRD, Rigaku D/Max2400), using
results from Newman et al. demonstrated that uncalcined Ru/ monochromatic Cu Ka radiation. The transmission electron
TiO2 catalysts with high dispersion of Ru metal exhibited high microscopy (TEM) measurements were performed on a JEOL-2010
selectivity to direct deoxygenation in HDO of phenol when transmission electron microscope. The size distribution of Pd particles
compared to Ru supported on carbon, alumina, and silica. They for every catalyst was estimated on the basis of 100 particles. X-ray
photoelectron spectroscopy (XPS) measurements were performed by
hypothesized that the outstanding catalytic activity of Ru/TiO2 using a Multilab2000 X-ray photoelectron spectrometer. The binding
was attributable to Ti3+ site created by hydrogen spillover, energies (BEs) data were calibrated by using the BE of C 1s at 284.8
interacting strongly with the oxygen in phenol to assist the eV as a reference. Prior to the characterization of XRD and XPS, the
breakage of C−O bond.41 Ardiyanti investigated bimetallic Ni− samples were treated by H2 (1 atm, 40 mL/min) at 200 °C for 3 h and
Cu catalysts on various supports for the catalytic hydrotreating then cooled to the room temperature followed passivation using 0.5%
of fast pyrolysis oil and reported that NiCu/TiO2 showed the O2/Ar mixture.
highest activity.42 In our previous work, the introduction of 2.3. Catalyst Activity Measurements. The evaluation test of
TiO2 to Pd/SiO2 catalyst significantly enhanced the HDO catalyst for HDO of guaiacol was carried out in a fixed-bed reactor
activity.43 with 6 mm inter diameter. In each run, 0.20 g sample (20−30 meshes)
was placed in the center of the reactor. Prior to feed the solution of 3
In the present work, we use guaiacol as a model compound wt % guaiacol in n-dodecane into the reactor, the palladium catalyst
to examine the activity of Pd supported on TiO2 with different was treated by H2 (1 atm, 40 mL/min) at 200 °C for 3 h and then
crystalline phase (anatase, rutile, and their mixture, P25) for the heated or cooled to the desired temperature. The HDO reaction were
HDO of guaiacol. For all the Pd/TiO2 catalysts, cyclohexane is performed at 200, 240, 260, and 280 °C, total pressure 2 MPa, and H2
the main deoxygenation product at elevated temperature, while flow rate 150 mL/min (at atmospheric pressure). After the reaction
2-methycyclohexanol is the only product on carbon supported was kept at the desired temperature for 6 h, the liquid product was
Pd catalysts. Pd supported on anatase TiO2 catalyst showed the collected by cooling the product mixture to room temperature and
highest HDO activity due to the more partially reduced then separating liquid phase from gas phase and was analyzed by off-
titanium species formation on the surface of anatase TiO2. line GC equipped with an FID detector and a commercial HP-5
column.
Guaiacol conversion (Xguaiacol), products selectivities (Sproducts‑i) were
2. EXPERIMENTAL SECTION calculated as follows: Xguaiacol = (Molguaiacol‑in − Molguaiacol‑out)/
2.1. Materials and Catalyst Preparation. PdCl2 solution (5 wt Molguaiacol‑in; Sproducts‑i = (Molproduct‑i × ni)/(Molreacted guaiacol × 7), ni
% in 10 wt % HCl), guaiacol (≥99.0%), and n-dodecane (≥99%) were refers to the carbon number in product-i.
bought from Sigma-Aldrich. Anatase TiO2, rutile TiO2, and Degussa
P25 TiO2 were bought from Jingrui New Material Corporation 3. RESULTS AND DISCUSSION
(Xuancheng, China)
Mesoporous carbon (MC) was prepared by using the method 3.1. Catalyst Characterization. The BET surface area,
reported in the literature.44 The properties of obtained MC are shown pore volume and average pore diameter of supports and as-
in Table 1. prepared catalysts are shown in Table 1. Generally, the BET
surface area and pore volume of TiO2 follows the order:
Table 1. Physical Properties of Supports and Catalysts TiO2(A) > TiO2(P) > TiO2(R).45 MC and AC have much
larger BET surface area than TiO2. The slight decrease in BET
surface area pore volume average pore diameter surface area and pore volume for all of catalysts after Pd
catalysts (m2/g) (m3/g) (nm)
deposition compared to the supports indicates that palladium
TiO2(A) 71 0.22 13.3 was deposited in the pores of the supports.
Pd/TiO2(A) 64 0.20 12.4 The TEM photographs of the as-prepared catalysts are
TiO2(P) 66 0.17 13.1 shown in Figure 1. It can be found that Pd particles are well
Pd/TiO2(P) 50 0.16 12.8 dispensed on the surface of supports. The average size of Pd
TiO2(R) 30 0.07 10.6 particles was estimated to be about 3.6 nm for Pd/TiO2(A),
Pd/TiO2(R) 26 0.06 9.6 which is slightly smaller than those for Pd/MC and Pd/AC
MC 626 0.81 4.5 according to histogram analysis, indicating that TiO2 is
Pd/MC 615 0.73 4.7 favorable for the dispersion of Pd particles.
AC 1297 1.23 2.7 The XRD patterns of all catalysts are shown in Figure 2. The
Pd/AC 1284 1.16 3.6
TiO2 structures were not altered by Pd loading and subsequent
thermal treatment. Crystalline Pd is hard to detect, except that
All catalysts were prepared with formaldehyde as reducing agent. In a small peak at 40.0° assigned to the Pd (111) plane46 can be
a typical preparation, the slurry of 2.0 g supports in 50 mL water was observed for Pd/TiO2(P), Pd/TiO2(R) and Pd/AC. Compared
added dropwise by 3.5 g of PdCl2 solution under stirring to achieve the
desired Pd loading of 5 wt %. Then the obtained suspension was to Pd/MC catalyst, Pd/AC catalyst exhibited a larger range of
stirred for 1 h. After the pH was adjusted to approximately 11 using a particle size, indicating that Pd is not well dispersed on the AC,
0.1 M NaOH solution, the solution was stirred for 2 h. Next, 1.0 mL which may be related to the pore texture of the AC support. It
formaldehyde (37 wt %) solution was slowly added to the suspension, is well know that the pores in AC are mainly micropores.
and the obtained suspension was stirred for an additional 60 min at 80 However, Pd nanoparticles are mainly in the internal surface of
10859 DOI: 10.1021/acs.energyfuels.7b01498
Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

MC due to its larger pores. Almost no XRD peak of Pd species


is observed for the Pd/TiO2(A) catalyst, suggesting that Pd
particles on TiO2(A) surface are smaller. The smaller Pd
particles obtained on TiO2(A) surface than on TiO2(P) and
TiO2(R) surface may be the result of larger surface area (Table
1). The results indicate that Pd particles are well dispersed on
supports, which are in good agreement with that obtained from
TEM.
The surface composition and oxidation states of all elements
(Pd, Ti, and C species) in catalysts were analyzed by using XPS.
Prior to the XPS test, the catalysts were treated at 200 °C,
followed passivation with 0.5% O2 in Ar flow after the catalysts
were cooled to room temperature. The spectra of Pd 3d, Ti 2p,
and C 1s were collected. The atom compositions of Pd, Ti, and
C on various catalysts surface are listed in Table 2.

Table 2. XPS Data of Catalysts


atomic composition
samples Pd Ti C Pd/Ti Pd/C
Pd/TiO2(A) 0.81 26.26 0.031
Pd/TiO2(P) 1.03 31.42 0.033
Pd/TiO2(R) 0.55 20.50 0.027
Pd/MC 0.33 90.17 0.0036
Pd/AC 0.74 85.49 0.0086

The Pd/Ti ratios in Pd/TiO2(A) and Pd/TiO2(P) catalysts


are close and higher than that in Pd/TiO2(R) catalyst,
indicating the properties of both Pd/TiO2(A) and Pd/
TiO2(P) catalysts are similar, which is different from Pd/
TiO2(R) catalyst. However, the Pd/Ti ratios for the three Pd/
TiO2 catalysts obtained from XPS analysis are lower than the
nominal value (0.04) for 5 wt % Pd/TiO2 catalyst, suggesting
that Pd nanoparticles not only located on the external surface of
TiO2 but in the pores of supports. For the Pd/MC catalyst,
however, Pd nanoparticles mainly locate on the internal surface
of MC due to its larger pore sizes; the Pd/C ratio obtained
from XPS for the Pd/MC catalysts is lower than the nominal
value (0.006). The reverse is true for Pd/AC. The surface
chemical states of Pd determined by XPS are shown in Figure
Figure 1. TEM images and particle size distribution of catalysts.
3. The core level XPS spectra of Ti 2p recorded from various
TiO2 samples (results not shown) show sharp and intense
peaks at 464.2 and 458.5 eV, which are assigned to Ti4+ in the
TiO2.47,48 However, No Ti3+ peaks were observed for all the

Figure 2. XRD profiles of catalysts (A anatase; R rutile).


Figure 3. XPS spectra of Pd 3d of the catalysts.

10860 DOI: 10.1021/acs.energyfuels.7b01498


Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

TiO2 samples under these conditions. It is known that binding Methoxycyclohexanol as the primary hydrogenation product at
energies of metallic Pd0 are at 334.9 ± 0.2 eV (3d5/2)49−51 low reaction temperature (200 °C) for all the Pd catalysts
while binding energies of 336.4 eV (3d5/2) and 337.6 eV confirms high hydrogenation ability of Pd catalysts. However,
(3d5/2) are assigned to Pd2+ and Pd4+ in PdO and PdO2, overwhelming amount of cyclohexane was found as the oxygen-
respectively.52,53 For Pd/TiO2 catalysts, a main peak located at free product on Pd/TiO2(A) and Pd/TiO2(P) while cyclo-
335 eV were observed companied a shoulder peak at 336.3 eV hexane was hardly found over Pd/TiO2(R), which suggests that
indicating that Pd species are mainly presented as Pd0 metal the breakage of C−O bond was facile on TiO2(A) and TiO2(P)
and a passivated layer of PdO are formed on the surface of Pd supported Pd catalysts. The selectivity of cyclohexane increased
nanoparticles. However, for the Pd/AC and Pd/MC catalysts, with the decrease in 2-methoxylcyclohexanol when the reaction
Pd species are mainly presented as PdO and PdO2 and no peak temperature was elevated. Especially, guaiacol almost trans-
for Pd0 metal is observed. The results suggest that the formed into cyclohexane and benzene completely at 260 °C
interaction between Pd and TiO2 occurs when Pd/TiO2 over Pd/TiO2(A) catalyst further indicating that Pd/TiO2(A)
catalysts were reduced at 200 °C. It is generally accepted that catalysts have high ability to improve the breakage of C−O
strong metal−support interaction between Pd and TiO2 bond. Benzene as oxygen-free product at elevated reaction
support occurs after high-temperature (above 500 °C) temperature over Pd/TiO2(A) catalysts was also found.
reduction, which will lead to the migration of partial reduced However, no benzene was formed during the hydrodeoxyge-
Ti3+ species onto the surface of noble metal particles.36 nation of cyclohexanol as the same conditions. Therefore, it is
However, Riyapan et al. found it was possible for the reduction likely that the direct breaking C−O bond of guaiacol rather
of Ti4+ to Ti3+ at relatively low reduction temperature in the than the dehydrogenation of cyclohexane leads to the
presence of Pd.54 The dissociatively chemisorbed hydrogen on formation of benzene, which suggests that a C−O direct
the Pd may spill over to TiO2 surface and reduce Ti4+ to breakage path exists besides the first hydrogenation and then
Ti3+.55−57 The formation of TiOx (x < 2) species at lower deoxygenation path in the HDO reaction on Pd/TiO2(A)
temperature was observed.58 Li et al. also reported that Ti4+ can catalyst at elevated reaction temperature.
be reduced to Ti3+ in the presence of Pd even at low reduction Pd supported on carbon catalysts, however, exhibited
temperature (200 °C).32 Partially reduced titanium species on dramatic difference in the products distribution compared to
the surface of TiO2 produced by the H2 reduction may be Pd supported on TiO2 catalysts. Guaiacol almost transformed
favorable to keep the palladium species in the form of metal to 2-methoxycyclohexanol completely, suggesting that carbon
Pd0. supported Pd catalysts have strong hydrogenation ability. A
3.2. Catalyst Activity. The guaiacol hydrodeoxygenation small amount of cyclohexanol and the almost absence of
was performed in the fixed-bed reactor at 200−280 °C under 3 oxygen-free products demonstrate that the hydrogenolysis
MPa. The conversions of guaiacol on all the catalysts are shown reaction of C−O bond is hard to occur on Pd/C catalysts likely
in Figure 4. It can be seen that all catalysts have high HDO because of the competing adsorption of aromatic ring with the
C−O bonds in the hydrogenation intermediate products.
Phenol and anisole, which are the main products for guaiacol
HDO on Pd/C catalysts,9,30 were not found over all the Pd
catalysts in this study, indicating that the catalytic HDO
performance of guaiacol was significantly influenced by the
catalysts employed and reaction conditions. The relative low
reaction temperature (<280 °C) and high pressure of H2
benefit the high selectivity of hydrogenation product 2-
mythoxylcyclohexanol.
It is very clear that the supports play a key role in the HDO
of guaiacol over supported Pd catalysts by analyzing the
product distribution. Over all the Pd catalysts, 2-mythoxylcy-
clohexanol as the main product at low reaction temperature
suggests that guaiacol was first hydrogenated on Pd particles. 2-
Figure 4. Conversion of guaiacol as a function of temperature on Methoxylcyclohexanol was further transformed to cyclohexanol
different catalysts.
and then to cyclohexane over Pd/TiO2 catalysts while no
further transformation of 2-methoxylcyclohexanol was found
activity of guaiacol at the temperature employed. Especially, over Pd supported on carbon. It is indicated that the adsorption
guaiacol is almost converted completely on all Pd catalysts of 2-methoxylcyclohexanol on Pd/TiO2 cataysts is substantially
except Pd/TiO2(R) suggesting that Pd catalysts have high different from that on Pd/carbon catalysts. The reduction of
hydrogenation activity. Compared to Pd/TiO2(R) catalyst, Pd/ Ti4+ to Ti3+ at low temperature (200 °C) was confirmed by
TiO2(A) and Pd/TiO2(P) catalysts exhibited higher hydro- EPR in the presence of Pd.32,55 Ti3+ on the surface of TiO2
genation, which may be due to the higher Pd content on the strongly interacts with the oxygen atom in the methanol and
catalyst surface (Table 2) and better dispersion of Pd metal results in the C−O bond scission.63,64 Therefore, we assume
(Figure 1). that the adsorption and breakage of C−O bond in 2-
The product distributions of guaiacol HDO over various methoxycyclohexanol mainly occur on the partially reduced
catalysts with temperature are shown in Figure 5. The guaiacol titanium species for Pd/TiO2 catalysts. Activated hydrogen
HDO over a number of catalysts were investigated.59−62 2- spillover from dissociated H2 on Pd to partially reduced
Methoxycyclohexanol, cyclohexanol, cyclohexane, and benzene titanium species react with chemisorbed C−O bond to give the
are found to be the main products over Pd/TiO2 catalysts products (Scheme 1). The amount of partially reduced titanium
within the temperature range examined (Figure 5a−c). 2- species on the surface of TiO2 decreased with the content of
10861 DOI: 10.1021/acs.energyfuels.7b01498
Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

Figure 5. Selectivity of products as a function of temperature over (a) Pd/TiO2(A), (b) Pd/TiO2(P), (c) Pd/TiO2(R), (d) Pd/MC, and (e)Pd/AC.

Scheme 1. Hydrodeoxygenation Reaction Pathways of To further investigate the active sites of Pd catalysts, the
Guaiacol on Pd/TiO2 and Pd/C Catalysts HDO of cyclohexanol over Pd/TiO2(A) and Pd/MC was
carried out. The conversion of cyclohexanol is shown in Figure
6. It is clear that the scission of C−O bond can take place on
both catalysts while the Pd/TiO2(A) exhibits higher C−O
bond scission activity. The particle sizes of Pd on the both
catalysts surface are comparable (Figures 1 and 2). Therefore,
partially reduced titanium species on the surface of TiO2 may

rutile in TiO2 due to its more stable structure.65 So the


difference of HDO activities of Pd supported on TiO2(A),
TiO2(P), and TiO2(R) catalysts could be attributed to the Figure 6. Conversion of cyclohexanol as a function of temperature
different amount of partially reduced titanium species. over catalysts.

10862 DOI: 10.1021/acs.energyfuels.7b01498


Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

play the substantial role in the C−O bond scission though Pd


metal also can catalyze this reaction. For Pd/MC and Pd/AC
catalysts, the hydrogenation of guaiacol and the C−O bond
breakage could occur on the same sites and the competing
adsorption of guaiacol and 2-methoxycyclohexanol may impede
the occurrence of C−O bond scission, which results in the
small amount of cyclohexane formed on Pd/MC and Pd/AC
catalysts. However, for Pd supported on TiO2 catalysts, the
hydrogenation of guaiacol takes place on Pd metal sites and
produced 2-methoxycyclohexanol diffuses to partially reduced
titanium species to react, which result in the high selectivity of
cyclohexane. This process is remarkably analogical to the case
of noble metal/acidic supports dual functional catalysts.27
The effect of H2 treating temperature of Pd/TiO2(A)
catalysts on the HDO property was investigated. The activity
and selectivity of Pd/TiO2(A) catalyst treated at different
temperatures are illustrated in Figures 7 and 8, respectively.

Figure 7. Conversion of guaiacol as a function of temperature on Pd/


TiO2(A) treated at different temperatures.

The Pd/TiO2(A) catalyst treated at 200 °C exhibited higher


activity than that treated at 50 °C. This may be due to the more
partially reduced titanium species generated by reduction at
higher temperature. Surprisingly although the activity of Pd/
TiO2(A) catalyst decreased dramatically, the selectivity of
Figure 8. Selectivity of products as a function of temperature over Pd/
cyclohexane increased when the catalyst was treated at higher
TiO2(A) treated by H2 at (a) 50, (b) 200, and (c) 500 °C.
temperature (500 °C) (Figure 8). The treatment at high
temperature does not alter the Pd particle size obviously
(Figure 9), which is in line with the results reported in the
literature.36 It is generally accepted that Pd surface will be
decorated by the migration of partially reduced TiOx when Pd/
TiO2 was treated by H2 at higher temperature (500 °C), which
will lead to the great decrease in the adsorption of H2.36
Therefore, the decrease in activity of Pd/TiO2(A) could be
attributed to the suppression of H2 chemisorption capacity of
Pd covered by the TiOx. And the higher ability to C−O bond
scission may result from more partially reduced titanium
species produced by treatment at high temperature. The results
further suggest that the partially reduced titanium species on
the surface of TiO2 contribute to the high activity of guaiacol
HDO over TiO2 supported Pd catalysts.

4. CONCLUSION
The HDO performance of Pd catalysts is substantially
influenced by the properties of supports. When carbon
materials, such as mesoporous carbon and activated carbon,
were employed as support of Pd catalysts for guaiacol HDO, 2- Figure 9. XRD profiles of Pd/TiO2(A) treated by H2 at different
methoxycyclohexanol as the dominating primary product was temperature (A anatase).
hardly transformed further. The C−O bond scission was
10863 DOI: 10.1021/acs.energyfuels.7b01498
Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

suppressed by the presence of guaiacol and 2-methoxycyclo- (11) Ferrari, M.; Bosmans, S.; Maggi, R.; Delmon, B.; Grange, P.
hexanol. The guaiacol HDO process on Pd/TiO2 catalysts was CoMo/carbon hydrodeoxygenation catalysts: influence of the hydro-
proposed that 2-methoxycyclohexanol generated by hydro- gen sulfide partial pressure and of the sulfidation temperature. Catal.
genation of guaiacol on Pd sites diffused to partially reduced Today 2001, 65 (2−4), 257−264.
(12) Bui, V. N.; Laurenti, D.; Afanasiev, P.; Geantet, C. Hydro-
titanium species and reacted with the active H atom from Pd to
deoxygenation of guaiacol with CoMo catalysts. Part I: Promoting
give cyclohexane. Pd supported on anatase TiO2 catalyst effect of cobalt on HDO selectivity and activity. Appl. Catal., B 2011,
showed that the highest HDO activity might be due to the 101 (3−4), 239−245.
more partially reduced titanium species formed on the surface (13) Ruiz, P. E.; Frederick, B. G.; De Sisto, W. J.; Austin, R. N.;
of anatase TiO2.


Radovic, L. R.; Leiva, K.; García, R.; Escalona, N.; Wheeler, M. C.
Guaiacol hydrodeoxygenation on MoS2 catalysts: Influence of
AUTHOR INFORMATION activated carbon supports. Catal. Commun. 2012, 27, 44−48.
Corresponding Authors (14) Mora, I. D.; Méndez, E.; Duarte, L. J.; Giraldo, S. A. Effect of
*Tel.: +86-519-86330360. Fax: +86-519-86330360. E-mail: support modifications for CoMo/γ-Al2O3 and CoMo/ASA catalysts in
the hydrodeoxygenation of guaiacol. Appl. Catal., A 2014, 474 (1),
mingshili@cczu.edu.cn (M.L.). 59−68.
*Tel.: +1-814-863-4466. E-mail: csong@psu.edu (C.S.). (15) Kubička, D.; Kaluža, L. Deoxygenation of vegetable oils over
ORCID sulfided Ni, Mo and NiMo catalysts. Appl. Catal., A 2010, 372, 199−
Mohong Lu: 0000-0002-3641-2619 208.
Chunshan Song: 0000-0003-2344-9911 (16) Senol, O. I.; Viljava, T. R.; Krause, A. O. I. Effect of sulphiding
agents on the hydrodeoxygenation of aliphatic esters on sulphided
Notes catalysts. Appl. Catal., A 2007, 326, 236−244.
The authors declare no competing financial interest.


(17) Bui, P.; Cecilia, J. A.; Oyama, S. T.; Takagaki, A.; Infantes-
Molina, A.; Zhao, H.; Li, D.; Rodríguez-Castellón, E.; Lopez, A. J.
ACKNOWLEDGMENTS Studies of the synthesis of transition metal phosphides and their
This work is financially supported by the National Natural activity in the hydrodeoxygenation of a biofuel model compound. J.
Catal. 2012, 294, 184−198.
Science Foundation of China (21761132006, 21676029). The
(18) Rensel, D. J.; Rouvimov, S.; Gin, M. E.; Hicks, J. C. Highly
financial support from Priority Academic Program Develop- selective bimetallic FeMoP catalyst for C−O bond cleavage of aryl
ment of Jiangsu Higher Education Institutions (PAPD) and the ethers. J. Catal. 2013, 305, 256−263.
EMS Energy Institute of the Pennsylvania State University at (19) Feitosa, L. F.; Berhault, G.; Laurenti, D.; Davies, T. E.; da Silva,
University Park are also acknowledged.


V. T. Synthesis and hydrodeoxygenation activity of Ni2P/C-Effect of
the palladium salt on lowering the nickel phosphide synthesis
REFERENCES temperature. J. Catal. 2016, 340, 154−165.
(1) Patel, M.; Kumar, A. Production of renewable diesel through the (20) Mortensen, P. M.; de Carvalho, H. W. P.; Grunwaldt, J. D.;
hydroprocessing of lignocellulosic biomass-derived bio-oil: A review. Jensen, P. A.; Jensen, A. D. Activity and stability of Mo2C/ZrO2 as
Renewable Sustainable Energy Rev. 2016, 58, 1293−1307. catalyst for hydrodeoxygenation of mixtures of phenol and 1-octanol. J.
(2) Demirbas, A. Biomass resource facilities and biomass conversion Catal. 2015, 328, 208−215.
processing for fuels and chemicals. Energy Convers. Manage. 2001, 42, (21) Lee, W. S.; Kumar, A.; Wang, Z.; Bhan, A. Chemical titration
1357−1378. and transient kinetic studies of site requirements in Mo2C-catalyzed
(3) van Haveren, J.; Scott, E. L.; Sanders, J. Bulk chemicals from vapor phase anisole hydrodeoxygenation. ACS Catal. 2015, 5, 4104−
biomass. Biofuels, Bioprod. Biorefin. 2008, 2, 41−57. 4114.
(4) Echeandia, S.; Pawelec, B.; Barrio, V. L.; Arias, P. L.; Cambra, J. (22) Zhang, W.; Zhang, Y.; Zhao, L.; Wei, W. Catalytic activities of
F.; Loricera, C. V.; Fierro, J. L. G. Enhancement of phenol NiMo carbide supported on SiO2 for the hydrodeoxygenation of ethyl
hydrodeoxygenation over Pd catalysts supported on mixed HY zeolite benzoate, acetone, and acetaldehyde. Energy Fuels 2010, 24, 2052−
and Al2O3. An approach to O-removal from bio-oils. Fuel 2014, 117, 2059.
1061−1073. (23) Ghampson, I. T.; Sepúlveda, C.; Garcia, R.; Fierro, J. L. G.;
(5) Boullosa-Eiras, S.; Lødeng, R.; Bergem, H.; Stocker, M.; Escalona, N.; DeSisto, W. J. Comparison of alumina- and SBA-15-
Hannevold, L.; Blekkan, E. A. Catalytic hydrodeoxygenation (HDO) supported molybdenum nitride catalysts for hydrodeoxygenation of
of phenol over supported molybdenum carbide, nitride, phosphide and guaiacol. Appl. Catal., A 2012, 435−436, 51−60.
oxide catalysts. Catal. Today 2014, 223, 44−53. (24) Ghampson, I. T.; Sepúlveda, C.; Garcia, R.; Frederick, B. G.;
(6) Jae, J.; Tompsett, G. A.; Foster, A. J.; Hammond, K. D.; Wheeler, M. C.; Escalona, N.; DeSisto, W. J. Guaiacol transformation
Auerbach, S. M.; Lobo, R. F.; Huber, G. W. Investigation into the over unsupported molybdenum-based nitride catalysts. Appl. Catal., A
shape selectivity of zeolite catalysts for biomass conversion. J. Catal. 2012, 413−414, 78−84.
2011, 279, 257−268. (25) Ghampson, I. T.; Sepúlveda, C.; Garcia, R.; Radovic, L. R.;
(7) Li, K.; Wang, R.; Chen, J. Hydrodeoxygenation of anisole over Fierro, J. L. G.; DeSisto, W. J.; Escalona, N. Hydrodeoxygenation of
silica-supported Ni2P, MoP, and NiMoP catalysts. Energy Fuels 2011, guaiacol over carbon-supported molybdenum nitride catalysts: Effects
25, 854−863. of nitriding methods and support properties. Appl. Catal., A 2012,
(8) Wildschut, J.; Iqbal, M.; Mahfud, F. H.; Cabrera, I. M.; 439−440, 111−124.
Venderbosch, R. H.; Heeres, H. J. Insights in the hydrotreatment of (26) Gutierrez, A.; Kaila, R. K.; Honkela, M. L.; Slioor, R.; Krause, A.
fast pyrolysis oil using a ruthenium on carbon catalyst. Energy Environ. O. I. Hydrodeoxygenation of guaiacol on noble metal catalysts. Catal.
Sci. 2010, 3, 962−970. Today 2009, 147 (3−4), 239−246.
(9) Shafaghat, H.; Rezaei, P. S.; Daud, W. M. A. W. Catalytic (27) Lee, C. R.; Yoon, J. S.; Suh, Y. W.; Choi, J. W.; Ha, J. M.; Suh,
hydrogenation of phenol, cresol and guaiacol over physically mixed D. J.; Park, Y. K. Catalytic roles of metals and supports on
catalysts of Pd/C and zeolite solid acids. RSC Adv. 2015, 5, 33990− hydrodeoxygenation of lignin monomer guaiacol. Catal. Commun.
33998. 2012, 17, 54−58.
(10) Laurent, E.; Delmon, B. Influence of water in the deactivation of (28) Lin, Y. C.; Li, C. L.; Wan, H. P.; Lee, H. T.; Liu, C. F. Catalytic
a sulfided NiMo/γ-Al2O3 catalyst during hydrodeoxygenation. J. Catal. Hydrodeoxygenation of guaiacol on Rh-based and sulfided CoMo and
1994, 146 (1), 281−291. NiMo catalysts. Energy Fuels 2011, 25, 890−896.

10864 DOI: 10.1021/acs.energyfuels.7b01498


Energy Fuels 2017, 31, 10858−10865
Energy & Fuels Article

(29) Zhao, C.; Kou, Y.; Lemonidou, A. A.; Li, X.; Lercher, J. A. (48) Price, N. J.; Reitz, J. B.; Madix, R. J.; Solomon, E. I. A
Highly selective catalytic conversion of phenolic bio-oil to alkanes. synchrotron XPS study of the vanadia−titania system as a model for
Angew. Chem. 2009, 121 (22), 4047−4050. monolayer oxide catalysts. J. Electron Spectrosc. Relat. Phenom. 1999,
(30) Sun, J.; Karim, A. M.; Zhang, H.; Kovarik, L.; Li, X. S.; Hensley, 98−99, 257−266.
A. J.; McEwen, J.; Wang, Y. Carbon-supported bimetallic Pd−Fe (49) Kaichev, V. V.; Morkel, M.; Unterhalt, H.; Prosvirin, I. P.;
catalysts for vapor-phase hydrodeoxygenation of guaiacol. J. Catal. Bukhtiyarov, V. I.; Rupprechter, G.; Freund, H. J. C−O bond scission
2013, 306, 47−57. on “defect-rich and perfect” Pd(1 1 1)? Surf. Sci. 2004, 566−568 (2),
(31) Zacher, A. H.; Olarte, M. V.; Santosa, D. M.; Elliott, D. C.; 1024−1029.
Jones, S. B. A review and perspective of recent bio-oil hydrotreating (50) Mamede, A. S.; Leclercq, G.; Payen, E.; Granger, P.;
research. Green Chem. 2014, 16, 491−515. Gengembre, L.; Grimblot, J. XPS characterization of adsorbed reaction
(32) Li, Y.; Fan, Y.; Yang, H.; Xu, B.; Feng, L.; Yang, M.; Chen, Y. intermediates on automotive exhaust gas catalysts: NO and CO + NO
Strong metal-support interaction and catalytic properties of anatase interactions with Pd. Surf. Interface Anal. 2002, 34, 105−111.
and rutile supported palladium catalyst Pd/TiO2. Chem. Phys. Lett. (51) Demoulin, O.; Navez, M.; Ruiz, P. The Activation of a Pd/γ-
2003, 372 (1−2), 160−165. alumina catalyst during methane combustion: Investigation of the
(33) Li, Y.; Xu, B.; Fan, Y.; Feng, N.; Qiu, A.; He, J. M.; Yang, H.; phenomenon and of potential causes. Catal. Lett. 2005, 103, 149−153.
Chen, Y. The effect of titania polymorph on the strong metal-support (52) Bi, Y.; Lu, G. Catalytic CO oxidation over palladium supported
interaction of Pd/TiO2 catalysts and their application in the liquid NaZSM-5 catalysts. Appl. Catal., B 2003, 41, 279−286.
phase selective hydrogenation of long chain alkadienes. J. Mol. Catal. (53) Ketteler, G.; Ogletree, D. F.; Bluhm, H.; Liu, H.; Hebenstreit, E.
A: Chem. 2004, 216 (1), 107−114. L. D.; Salmeron, M. In situ spectroscopic study of the oxidation and
(34) Koussathana, M.; Vamvouka, N.; Tsapatsis, M.; Verykios, X. reduction of Pd(111). J. Am. Chem. Soc. 2005, 127, 18269−18273.
Hydrogenation of aromatic compounds over noble metals dispersed (54) Riyapan, S.; Boonyongmaneerat, Y.; Mekasuwandumrong, O.;
on doped titania carriers. Appl. Catal., A 1992, 80 (1−2), 99−113. Praserthdam, P.; Panpranot, J. Effect of surface Ti3+ on the sol−gel
(35) Binder, A.; Seipenbusch, M.; Muhler, M.; Kasper, G. Kinetics derived TiO2 in the selective acetylene hydrogenation on Pd/TiO2
and particle size effects in ethene hydrogenation over supported catalysts. Catal. Today 2015, 245, 134−138.
palladium catalysts at atmospheric pressure. J. Catal. 2009, 268 (1), (55) Xu, J.; Sun, K.; Zhang, L.; Ren, Y.; Xu, X. A highly efficient and
150−155. selective catalyst for liquid phase hydrogenation of maleic anhydride to
(36) Fu, Q.; Wagner, T.; Olliges, S.; Carstanjen, H. Metal−oxide butyric acid. Catal. Commun. 2005, 6, 462−465.
interfacial reactions: Encapsulation of Pd on TiO2 (110). J. Phys. Chem. (56) Huizinga, T.; Prins, R. Behavior of titanium(3+) centers in the
B 2005, 109 (2), 944−951. low- and high-temperature reduction of platinum/titanium dioxide,
(37) Weerachawanasak, P.; Mekasuwandumrong, O.; Arai, M.; Fujita, studied by ESR. J. Phys. Chem. 1981, 85, 2156−2158.
S.; Praserthdam, P.; Panpranot, J. Effect of strong metal−support (57) Conesa, J. C.; Soria, J. Reversible titanium(3+) formation by
interaction on the catalytic performance of Pd/TiO2 in the liquid- hydrogen adsorption on M/anatase (TiO2) catalysts. J. Phys. Chem.
phase semihydrogenation of phenylacetylene. J. Catal. 2009, 262 (2), 1982, 86, 1392−1395.
199−205. (58) Panagiotopoulou, P.; Christodoulakis, A.; Kondarides, D. I.;
(38) Tapin, B.; Epron, F.; Especel, C.; Ly, B. K.; Pinel, C.; Besson, M. Boghosian, S. Particle size effects on the reducibility of titanium
Study of Monometallic Pd/TiO2 Catalysts for the hydrogenation of dioxide and its relation to the water−gas shift activity of Pt/TiO2
succinic acid in aqueous phase. ACS Catal. 2013, 3 (10), 2327−2335. catalysts. J. Catal. 2006, 240, 114−125.
(39) Boonyasuwat, S.; Omotoso, T.; Resasco, D. E.; Crossley, S. P. (59) Gao, D.; Schweitzer, C.; Hwang, H. T.; Varma, A. Conversion of
Conversion of guaiacol over supported Ru catalysts. Catal. Lett. 2013, guaiacol on noble metal catalysts: Reaction performance and
143, 783−791. deactivation studies. Ind. Eng. Chem. Res. 2014, 53 (49), 18658−
(40) Omotoso, T.; Boonyasuwat, S.; Crossley, S. P. Understanding 18667.
the role of TiO2 crystal structure on the enhanced activity and stability (60) Lee, K.; Gu, G. H.; Mullen, C. A.; Boateng, A. A.; Vlachos, D. G.
of Ru/TiO2 catalysts for the conversion of lignin-derived oxygenates. Guaiacol hydrodeoxygenation mechanism on Pt(111): Insights from
Green Chem. 2014, 16, 645−652. density functional theory and linear free energy relations. Chem-
(41) Newman, C.; Zhou, X.; Goundie, B.; Ghampson, I. T.; Pollock, SusChem 2015, 8 (2), 315−322.
R. A.; Ross, Z.; Wheeler, M. C.; Meulenberg, R.; Austin, R. N.; (61) Pinheiro, A.; Hudebine, D.; Dupassieux, N.; Geantet, C. Impact
Frederick, B. G. Effects of support identity and metal dispersion in of oxygenated compounds from lignocellulosic biomass pyrolysis oils
supported ruthenium hydrodeoxygenation catalysts. Appl. Catal., A on gas oil hydrotreatment. Energy Fuels 2009, 23 (2), 1007−1014.
2014, 477, 64−74. (62) Yan, N.; Yuan, Y.; Dykeman, R.; Kou, Y.; Dyson, P. J.
(42) Ardiyanti, A. R.; Khromova, S. A.; Venderbosch, R. H.; Hydrodeoxygenation of lignin-derived phenols into alkanes by using
Yakovlev, V. A.; Melian-Cabrera, I. V.; Heeres, H. J. Catalytic nanoparticle catalysts combined with brønsted acidic ionic liquids.
Angew. Chem., Int. Ed. 2010, 49 (32), 5549−5553.
hydrotreatment of fast pyrolysis oil using bimetallic Ni−Cu catalysts
(63) Zhang, Z.; Bondarchuk, O.; White, J. M.; Kay, B. D.; Dohnálek,
on various supports. Appl. Catal., A 2012, 449, 121−130.
Z. Imaging adsorbate O−H bond cleavage: methanol on TiO2(110). J.
(43) Lu, M. H.; Zhu, J.; Li, M. S.; Shan, Y. H.; He, M. Y.; Song, C. S.
Am. Chem. Soc. 2006, 128, 4198−4199.
TiO2-modified Pd/SiO2 for catalytic hydrodeoxygenation of guaiacol.
(64) Henderson, M. A.; Otero-Tapia, S.; Castro, M. E. The chemistry
Energy Fuels 2016, 30, 6671−6676.
of methanol on the TiO2(110) surface: the influence of vacancies and
(44) Huang, Y.; Hu, S.; Zuo, S.; Xu, Z.; Han, C.; Shen, J. Mesoporous
coadsorbed species. Faraday Discuss. 1999, 114, 313−329.
carbon materials prepared from carbohydrates with a metal chloride
(65) Panpranot, J.; Kontapakdee, K.; Praserthdam, P. Effect of TiO2
template. J. Mater. Chem. 2009, 19, 7759−7764.
crystalline phase composition on the physicochemical and catalytic
(45) Ohno, T.; Sarukawa, K.; Matsumura, M. Photocatalytic activities
properties of Pd/TiO2 in selective acetylene hydrogenation. J. Phys.
of pure rutile particles isolated from TiO2 powder by dissolving the
Chem. B 2006, 110, 8019−8024.
anatase component in HF solution. J. Phys. Chem. B 2001, 105, 2417−
2420.
(46) Wang, X.; Xia, Y. The influence of the crystal structure of TiO2
support material on Pd catalysts for formic acid electrooxidation.
Electrochim. Acta 2010, 55, 851−856.
(47) Jing, L.; Sun, X.; Cai, W.; Xu, Z.; Du, Y.; Fu, H. The preparation
and characterization of nanoparticle TiO2/Ti films and their
photocatalytic activity. J. Phys. Chem. Solids 2003, 64, 615−623.

10865 DOI: 10.1021/acs.energyfuels.7b01498


Energy Fuels 2017, 31, 10858−10865

Você também pode gostar