You are on page 1of 5

Catalysis Communications 94 (2017) 18–22

Contents lists available at ScienceDirect

Catalysis Communications

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

Short communication

Effect of Ti loading on structure-activity properties of Cu-Ni/Ti-MCM-41


catalysts in hydrodeoxygenation of guaiacol
S.B. Abd Hamid a,⁎, Murtala M. Ambursa a,b, Putla Sudarsanam c, Lee Hwei Voon a, Suresh K. Bhargava c
a
Nanotechnology and Catalysis Research Centre (NANOCAT), University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Department of Chemistry, Kebbi State University of Science and Technology Aliero, Nigeria
c
Centre for Advanced Materials and Industrial Chemistry (CAMIC), School of Science, RMIT University, Melbourne, Victoria 3001, Australia

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

Article history: Various Cu-Ni/Ti-MCM-41 catalysts by varying Ti amount (Ti/Si = 10, 20 and 30%) were synthesized for
Received 8 November 2016 hydrodeoxygenation of guaiacol and characterized by XRD, N2 adsorption-desorption, NH3-TPD, H2-TPR and
Received in revised form 5 February 2017 Raman spectroscopy. The Ti loading shows an adverse effect on the catalyst surface area. But, more number of
Accepted 6 February 2017
acidic sites (5173.55 μmol/g) were found for 20% Ti loaded Cu-Ni/Ti-MCM-41 catalyst. The 20% Ti loaded Cu-
Available online 8 February 2017
Ni/Ti-MCM-41 catalyst shows a high guaiacol conversion (74.2%) and cyclohexane selectivity (48.81%) which
Keywords:
is due to abundant acid and redox sites: acid sites initiate removal of oxygen via dehydration, while hydrogena-
Biomass valorisation tion of guaiacol to cyclohexane is promoted by redox sites.
Lignin model compound © 2017 Published by Elsevier B.V.
Guaiacol
Hydrodeoxygenation
Cu-Ni/Ti-MCM-41
Ti loading

1. Introduction Guaiacol molecule has been used as lignin model compound [6] be-
cause it contains two types of O-containing functional groups (Caryl-
Currently, the transformation of renewable biomass feed stocks into OCH3 and Caryl-OH), which are characteristic of lignin fragment [7,8].
valuable chemicals and fuels have gained considerable attention due to It has been demonstrated that catalysts having both redox and acid
continuous depletion of non-renewable fuel sources. Lignocellulose is sites play a key role in the removal of oxygen from lignin via HDO pro-
the second most abundant biomass feedstock after the marine micro/ cess [9]. Consequently, various bi-functional catalysts composed of
macro-algae biomass and does not compete with food supplies because noble metals (Ru, Pt, Pd, and Re) with different supports, namely TiO₂,
of its non-edible nature [1,2]. Lignocellulose is mainly composed of cel- CeO₂, and SiO₂ have been developed for the HDO of lignin model com-
lulose, hemicellulose, and lignin. Compared to cellulose and hemicellu- pounds [7,8,10–20]. However, the catalytic applications of noble metals
lose, lignin valorisation remains a major challenge because of its are limited due to their high cost and low abundance. Other studies
complex polyaromatic structure [3]. Generally, lignin is directly con- using non-noble metals, such as Ni, NiMo, Mo, Fe, Ni2P, Co2P, etc. sup-
verted into bio-oils through pyrolysis and liquefaction methods but ported on acidic supports such as Al2O3, SiO2, ZrO₂, [21–25] have been
the pyrolysis method tend to be more promising. However, this process also explored. As well, metal oxide supports show smaller surface
exhibit large fraction of lignin-derived compounds with high amounts area, leading into aggregation of active phases, hence low catalytic
of oxygen and oxo-functionalized groups that lower their energy densi- performance.
ty relative to conventional hydrocarbon fuels [4]. Catalytic In this context, the combination of non-noble metals like Cu and Ni
hydrodeoxygenation (HDO) of lignin using molecular hydrogen is with large surface area mesoporous silica materials seems to be best
seemed to be more advantageous than other conversion processes choice of catalytic system for HDO reaction of guaiacol. Both Cu and Ni
since, it produces oil with little amount of oxygen that can be easily are cheap, high abundant and exhibit superior redox properties. As
upgraded to hydrocarbon fuels [4]. Due to complex nature of lignin frag- well, mesoporous silica materials, for example MCM-41 shows a high
ment, the use of model compounds in hydrodeoxygenation processes surface area and adequate acidic sites. It has been demonstrated that
has been adopted [5]. the properties of MCM-41 material can be improved by incorporating
heteroatoms such as Ti [26,27]. Moreover, Ti species play a key role in
deoxygenation reaction [28]. It is therefore expected that the dispersion
⁎ Corresponding author. of bimetallic CuNi particles on the surface of Ti-MCM-41 may provide
E-mail address: sharifahbee.ahamid@nanoc.com.my (S.B. Abd Hamid). adequate redox and acid properties for efficient HDO of guaiacol.

http://dx.doi.org/10.1016/j.catcom.2017.02.006
1566-7367/© 2017 Published by Elsevier B.V.
S.B. Abd Hamid et al. / Catalysis Communications 94 (2017) 18–22 19

Therefore, this work investigates the synthesis of Cu-Ni/Ti-MCM-41 cat- Nr.20022896 with a thermal conductivity detector (TCD). For each ex-
alysts by varying the Ti amount (Ti/Si = 10, 20 and 30%) for HDO of periment, 51.30 mg of the catalyst was first outgassed at 120 °C for
guaiacol. A systematic characterization has been undertaken to eluci- 1 h in He gas, followed by cooling the system to 50 °C before following
date the influence of Ti on the acid, redox, and catalytic properties of 10% NH3/Heat 20 ml/min for 1 h. The weakly physisorbed NH3 was re-
NiCu/Ti-MCM-41 catalysts. moved by purging with 100% helium gas for I hr. Then, thermal desorp-
tion of NH3 was carried out between the temperature of 100–900 °C at
10 °C/min in a He atmosphere flowing at 20 ml/min with a final holding
2. Experimental
time of 30 min. Similarly, H2-TPR was conducted on the same instru-
ment. In a typical procedure, 50 mg of the sample was heated from 30
2.1. Catalysts synthesis
to 650 °C at a heating rate of 10 °C/min with 20% H2 in N2 at a flow
rate of 40 ml/min with a terminal holding time of 60 min. Raman anal-
Various Ti-MCM-41 materials with different Ti/Si atomic ratios (10/
ysis was conducted using Renishaw in via confocal Raman microscope
90, 20/80 and 30/70) were prepared by modifying the reported proce-
having charge coupled device with visible laser of 514 nm.
dure [29]. The detailed synthesis procedure was presented in Supple-
mentary information (ESI). The NiCu (10 wt% as 3:1 of Ni:Cu)
catalysts supported on Ti-MCM-41 samples were synthesized according 2.3. Catalysts performance test
to the procedure reported elsewhere [7]. 1.06 g Ni(NO3)2·6H2O (99%,
Aldrich) and 0.48 g Cu(NO3)2·3H2O (99%, Aldrich) were dissolved in Prior to the reaction, the catalysts were reduced in 20% H2 in He with
deionised water. This solution was added to another solution containing a flow rate of 40 ml/min at 550 °C for 5 h. Then, 0.1 g of the reduced cat-
4.25 g of Ti-MCM-41 in deionised water and aged at 70 °C with stirring alyst (inside glass bulb), 96 wt% heptane, and 4 wt% guaiacol were
for 6 h. The resulting viscous solution was dried at 90 °C for 18 h and transferred into the reactor. The reactions were performed at 10 MPa
then calcined in air at 550 °C for 4 h. H2 pressure, 260 °C, 1000 rpm and 6 h. The reaction products were
analysed qualitatively using GC–MS (Agilent GC-MS6890N) with HP-
5MS (30 m × 0.250 mm × 0.25 μm) capillary column and quantitatively
2.2. Catalysts characterization
using GC-FID (Agilent GC-6890N) equipped with HP-5
(30 m × 0.53 mm).
The small angle XRD analysis was carried out using small angle X-ray
scattering diffractometer (Bruker) with Cu Kα (λ = 0.1542 nm) in the
range of 0 to 10° with step size of 0.01° and step time of 5 s. The wide 3. Results and discussion
angle XRD analysis was performed using X-ray diffractometer (Bruker
D8 advance) with Cu Kα (λ = 0.1542 nm) radiation run at 40 kV and 3.1. Physicochemical characterization
40 mA in the range of 2θ = 2 to 80° with the step size of 0.03° and
step time of 1 s. Then the N2 adsorption-desorption analysis was con- Small angle XRD patterns of Ti-MCM-41 samples show a high inten-
ducted (Tristar II series; micrometric) to estimate the textural proper- sity peak at (100) plane (Fig. S1, ESI), indicating the formation of or-
ties of the samples. Prior to analysis, 300 mg of the sample was dered hexagonal cylindrical structures as a model for MCM-41 [30].
outgassed at 300 °C for 5 h in the analyser sample tube. The NH3-TPD With the increase of Ti loading from 20 to 30%, there was a peak shifting
analysis was conducted on TPDRO 1100 instrument with Ser. (from 2.85 to 2.70°) toward lower angle side. This is due to the

Fig. 1. NH3-TPD profiles of Ti-MCM-41 samples with different Ti loadings.


20 S.B. Abd Hamid et al. / Catalysis Communications 94 (2017) 18–22

distortion of hexagonal structure of MCM-41, an evidence for the incor-


poration of titanium species within the silica matrix [31]. Owing to
higher ionic radius of titanium (Ti4+ = 0.072 nm) [32] compared to sil-
icon (Si4+ = 0.041 nm) [33], the bond length in Ti-O-Si of Ti-MCM-41
could be larger than in Si-O-Si of pure MCM-41, resulting in the distor-
tion of hexagonal structure of MCM-41.
N2 adsorption-desorption isotherms of Ti-MCM-41 samples show a
type IV with H3 hysteresis loop, indicating mesoporous nature of the
materials (Fig. S2, ESI) [34]. The BET surface area of Ti-MCM-41 sample
is decreased from ~833, 731, and 496 m2/g with the Ti loading from 10,
20 and 30%, respectively (Table S1, ESI). Similarly, the pore volume of Ti-
MCM-41 sample is decreased from ~0.5, ~0.4, and ~0.3 cm3/g with the
Ti loading from 10, 20 and 30%, respectively. The decrease of BET surface
area and pore volume at higher Ti loadings could be due to collapsed of
other part within the MCM-41 structure as evidenced by XRD studies
(Fig. S1, ESI) as Ti is heavier than Si. On the contrary, the pore size of Fig. 2. Guaiacol conversion over Cu-Ni/Ti-MCM-41 catalysts at 260 °C, 10 MPa H2 pressure
and 6 h.
Ti-MCM-41 increases from ~2.6, ~2.7, and ~3.0 nm with the Ti loading
from 10, 20, and 30%, respectively which is due to more Ti coordination
in the silicate structures.
The NH3-TPD profiles of Ti-MCM-41 samples are presented in Fig. 1. ~24° is a manifestation of amorphous structures in the synthesized ma-
The peak at low temperature (394 °C) indicates desorption of NH3 from terials [47].
the weak acid sites [35–37] which is shifted to lower temperatures with All the Cu-Ni/Ti-MCM-41 catalysts exhibit type IV isotherm with H3
the increase of Ti loading. This could be affiliated to poor distribution of hysteresis loop (Fig. S6, ESI). The BET surface area of Ti-MCM-41 sam-
coordinated titanium species in MCM-41 frame work. On the contrary, ples were found to decrease from ~833, ~731, and ~497 m2/g to ~512,
the high temperature peak at 555 °C (desorption of NH3 from strong ~497, and ~205 m2/g after the addition of Cu and Ni to 10, 20 and 30%
acidic sites) is shifted to high temperatures with the increase of Ti load- Ti-MCM-41 samples, respectively (Tables S1 and S2). Also, the pore vol-
ing. These observations indicate that the strength of strong acidic sites ume and average pore size were found to decrease to ~0.3–~0.19 cm3/g
increases as the titanium loading increases from 10, 20 and 30%. The and ~ 2.5–~2.3 nm after the addition of Cu and Ni to Ti-MCM-41 sam-
data presented in Table S1, ESI reveal that the 20% Ti loaded MCM-41 ples, respectively. The decreased textural properties in supported cata-
sample has high concentration of acidic sites, which could play a lysts could be due to the occupation of dispersed CuO and NiO species
favourable role in HDO of guaiacol as discussed in the later paragraphs. within the internal and external surface of Ti-MCM-41 supports.
The H2-TPR profiles of Cu-Ni/Ti-MCM-41 samples are shown in
(Fig. S3, ESI). The position of reduction peaks of both copper and nickel
oxides is shifted to higher temperatures with the increase of Ti loading. 3.2. Hydrodeoxygenation of guaiacol over CuNi/Ti-MCM-41 catalysts
This indicates an enhancement in the interactions of copper and nickel
oxides with Ti-MCM-41 as Ti loading increases. The observed peaks at The obtained results in hydrodeoxygenation of guaiacol over CuNi/
240 and 280 °C for 10 and 20% Ti loaded samples, respectively, could Ti-MCM-41 catalysts are shown in Fig. 2. It was found that, the conver-
be assigned to reduction of CuO to Cu, while the peaks noticed at 330 sion of guaiacol increases from 51.5 to 74.2%, when the Ti loading in-
and 353 °C are associated with the reduction of NiO to Ni [38,39]. The creases from 10 to 20% and then decreases to 55.9% for 30% Ti loaded
30% Ti loaded sample shows various peaks; the initial tail at 250 °C rep- catalyst. Fig. 3 shows selectivity of cyclohexane obtained in HDO of
resents the reduction of highly exposed and less interacted CuO to Cu guaiacol over NiCu/Ti-MCM-41 catalysts. The selectivity of cyclohexane
metal. The bands at 310 and 390 °C could be attributed to the reduction was found to increase from 37.7 to 48.8% when the Ti loading increases
of bulk CuO to Cu and NiO to Ni, respectively [40]. from 10 to 20% in NiCu/Ti-MCM-41 catalyst. In contrast, the cyclohex-
Two major Raman bands are noticed for 10 and 20% Ti loaded sam- ane selectivity decreases from 48.8 to 39.0% with the increase of Ti load-
ples (Fig. S4, ESI): the first peak located at 580 cm− 1 is assigned to ing from 20 to 30%, respectively. The observed high conversion of
Ni\\O symmetric stretching vibration while the second peak at guaiacol and high yield of cyclohexane in the case of 20% Ti loaded
1077 cm− 1 indicates the asymmetric stretching vibration mode of NiCu/Ti-MCM-41 catalyst is due to well-balanced acid and redox sites.
Ni\\O [41,42]. No Raman peaks of CuO were found for 10 and 20% Ti
loaded samples, which could be attributed to low Cu loading. The no-
ticed peaks at ~576 and 1060 cm−1 for 30% Ti loaded catalyst is ascribed
to symmetric and asymmetric stretching modes of Ni\\O, respectively
[43]. The peak at ~690 cm−1 is assigned to vibrational mode of Cu\\O.
A red shifting from 580 and 1077 cm−1 to 576 and 1060 cm−1 is noticed
for 30% Ti loaded sample, respectively, which could be related to high
degree of interaction between NiO and the support, in line with H2-
TPR results (Fig. S3, ESI).
The XRD patterns (Fig. S5, ESI) of Ni-Cu/Ti-MCM-41 samples show
various peaks at ~38°, 43° and 63°, corresponding to (111), (200) and
(220) planes of cubic phase NiO, respectively [44,45]. The 30% Ti loaded
Cu-Ni/Ti-MCM-41 catalyst shows two peaks at ~ 36 and ~ 39°, corre-
sponding to (002) and (111) planes of CuO, respectively [46]. But,
only (002) plane of CuO was noticed for 20% Ti containing Cu-Ni/Ti-
MCM-41, while no CuO XRD peaks were observed for 10% Ti containing
Cu-Ni/Ti-MCM-41. The observed small surface area in 30% Ti containing
MCM-41 sample (Table S1, ESI) could be attributed to larger particles Fig. 3. Selectivity of cyclohexane in HDO of guaiacol over NiCu/Ti-MCM-41 catalysts at
which could be detected by XRD analysis. The noticed broad peak at 260 °C, 10 MPa H2 pressure and 6 h.
S.B. Abd Hamid et al. / Catalysis Communications 94 (2017) 18–22 21

Table 1 hydrocarbons products. The total distribution of pure hydrocarbons


Products distribution from HDO of guaiacol over Ni-Cu/Ti-MCM-41 catalysts with different were found to increase in the following order 20% (80.2%) N 10%
Ti loading from 10, 20 and 30%.
(59.3) N 30% (43.3%) Ti loaded NiCu/Ti-MCM-41 catalysts. Based on
Products Products distributions (%) this distribution, along with the reported literature articles [48–50],
10% 20% 30% the possible reaction pathways were proposed (Scheme 1). According
to Scheme 1, HDO of guaiacol proceeds via dihydroxylation and
Cyclohexane 37.7 48.8 39.0
Cyclohexene 15.5 11.2 4.3 demethoxylation pathways over acid sites to yield phenol and anisole
Cyclohexanol 12.8 15.2 14.6 and then metal sites promote the hydrogenation of aryl ring to yield cy-
Cyclohexanone 7.1 4.2 16.2 clohexanone and methyl-cyclohexane. Subsequent hydrogenation, de-
Phenol 20.8 – 15.4 hydration and further hydrogenation convert cyclohexanone into
Methyl cyclohexane 4.1 19.0 –
Bicyclohexyl 2.0 1.1 –
cyclohexanol, cyclohexene and cyclohexane, respectively. On the other
Anisole – – 1.8 hand, methoxy-cyclohexane can either totally be deoxygenated and in-
Methoxycyclohexane – – 8.7 stantly methylated by methyl transfer to form methyl cyclohexane or
can produce cyclohexane which transform into bicyclohexyl through
ring coupling reactions. Further studies are underway to optimise the
The number of metal sites is expected to be the same due to similar reaction conditions for achieving maximum yield of cyclohexane over
metal loading of Cu and Ni in all the Cu-Ni/Ti-MCM-41 catalysts. But, 20% Ti loaded NiCu/Ti-MCM-41 catalyst.
the number of acids sites varied with a high concentration of these
acids sites over 20% Ti loaded NiCu/Ti-MCM-41 catalyst (Fig. 1 and
Table S1, ESI). Acids sites play a major role in the oxygen removal 4. Conclusions
through dehydration, whereas the hydrogenation of guaiacol to yield
cyclohexane is promoted by the redox sites [47]. In summary, a series of Ti-MCM-41 supports with various Ti loadings
Table 1 shows products distribution in HDO of guaiacol over NiCu/Ti- (Ti/Si ratio = 10, 20 and 30%) and the corresponding supported Cu\\Ni
MCM-41 catalysts. In addition to cyclohexane as a major product, eight catalysts were prepared and fully characterized their structural, acidic,
other different products including cyclohexene, cyclohexanol, cyclohex- and redox properties. The investigated characterization results showed
anone, methyl-cyclohexane, bicyclohexyl, anisole, methoxy- that the increase of Ti loading decrease the surface area of the catalyst.
cyclohexane and phenol were found over NiCu/Ti-MCM-41 catalysts. But, the concentration of acid sites increases and decreases with Ti load-
Anisole and methoxycyclohexane were not observed over 10% Ti loaded ing having the highest concentration of acidic sites on 20% Ti loaded
CuNi catalyst (Table 1). In addition to these two products, phenol was CuNi/Ti-MCM-41 catalyst.
also absent over 20% Ti loaded NiCu/Ti-MCM-41 catalyst. There are The performance of Cu-Ni/Ti-MCM-41 catalysts for HDO of guaiacol
four pure hydrocarbons found including cyclohexane, cyclohexene, shows that, guaiacol conversion and selectivity of cyclohexane were ob-
methyl-cyclohexane and bicyclohexyl over 10% and 20% Ti loaded served to increase and decrease with increased of Ti loading. The 20% Ti
NiCu/Ti-MCM-41 catalysts. On the other hand, methyl-cyclohexane loaded NiCu/Ti-MCM-41 shows a best performance in terms of guaiacol
and bicyclohexyl were not found over 30% Ti loaded NiCu/Ti-MCM-41 conversion (74.2%) and cyclohexane selectivity (48.8%). The presence of
catalysts. This explains that, as Ti loading increases, products selectivity ample amount of acidic and redox sites are found to be key factors for
changes toward oxygenated products. The possible reason is that, the high performance of 20% Ti loaded NiCu/Ti-MCM-41 catalyst in
number of acid sites generated by coordinated Ti species beyond 20% hydrodeoxygenation of guaiacol.
Ti loading is decreased and poorly dispersed due to decrease in the sur-
face area of the catalysts. The decrease in surface area could result to
poor distribution of coordinated species leading to strength loss of Acknowledgement
acids sites to weak ones. This type of behaviour could reduce deoxygen-
ation performance of the catalysts leading to more oxygenated products The authors are grateful for scholarship from Kebbi State University
as observed from 20% to 30% Ti loaded catalysts. Similar to this observa- of Science and Technology Aliero, Nigeria and funding from University
tion in which textural properties and acidity of the catalysts strongly in- of Malaya, Malaysia (Grand challenge: GC grant no: GC001A-14AET).
fluenced HDO performance has been reported elsewhere; the catalysts
with higher acidity and better textural properties show better perfor- Appendix A. Supplementary data
mance in HDO reactions [17] similar observation to 20% Ti loaded
NiCu/Ti-MCM-41 catalysts. The cooperative role of metal functions on Supplementary data to this article can be found online at http://dx.
deoxygenated compounds converts unsaturated to saturated doi.org/10.1016/j.catcom.2017.02.006.

Scheme 1. Possible reactions pathways for Guaiacol HDO over Cu\


\Ni supported on Ti-MCM-41.
22 S.B. Abd Hamid et al. / Catalysis Communications 94 (2017) 18–22

References [26] D.P. Sahoo, D. Rath, B. Nanda, K. Parida, Transition metal/metal oxide modified
MCM-41 for pollutant degradation and hydrogen energy production: a review,
RSC Adv. 5 (102) (2015) 83707–83724.
[1] S.C. Phulara, P. Chaturvedi, P. Gupta, Isoprenoid-based biofuels: homologous and
[27] B. Jiang, L. Wang, Y. Sun, Y. Ma, RSC Adv., (technology 10 14).
heterologous expressions in prokaryotes, Appl. Environ. Microbiol. (2016) (AEM.
[28] M. Selvaraj, K. Shanthi, R. Maheswari, A. Ramanathan, Hydrodeoxygenation of
01192-16).
Guaiacol over MoO3-NiO/mesoporous silicates: effect of incorporated heteroatom,
[2] L. Wang, H. Wan, S. Jin, X. Chen, C. Li, C. Liang, Hydrodeoxygenation of dibenzofuran
Energy Fuel 28 (4) (2014) 2598–2607.
over SiO2, Al2O3/SiO2 and ZrO2/SiO2 supported Pt catalysts, Catal. Sci.Tech. 5 (1)
[29] B. Boukoussa, S. Zeghada, G.B. Ababsa, R. Hamacha, A. Derdour, A. Bengueddach, F.
(2015) 465–474.
Mongin, Catalytic behavior of surfactant-containing-MCM-41 mesoporous materials
[3] S. Dutta, K.C.-W. Wu, B. Saha, Emerging strategies for breaking the 3D amorphous
for cycloaddition of 4-nitrophenyl azide, Appl. Catal. A Gen. 489 (2015) 131–139.
network of lignin, Catal. Sci. Tech. 4 (11) (2014) 3785–3799.
[30] B. Marler, U. Oberhagemann, S. Vortmann, H. Gies, Influence of the sorbate type on
[4] M. Saidi, F. Samimi, D. Karimipourfard, T. Nimmanwudipong, B.C. Gates, M.R.
the XRD peak intensities of loaded MCM-41, Microporous Mater. 6 (5) (1996)
Rahimpour, Upgrading of lignin-derived bio-oils by catalytic hydrodeoxygenation,
375–383.
Energy Environ. Sci. 7 (1) (2014) 103–129.
[31] G.A. Eimer, S.G. Casuscelli, C.M. Chanquia, V. Elías, M.E. Crivello, E.R. Herrero, The in-
[5] S. Sarkar, Experimental & Computational Approaches Towards Improving Mesoscale
fluence of Ti-loading on the acid behavior and on the catalytic efficiency of mesopo-
Understanding of Particle Interactions in Pharmaceutical Systems, 2015.
rous Ti-MCM-41 molecular sieves, Catal. Today 133 (2008) 639–646.
[6] X. Tan, X. Zhuang, S. Lü, W. Qi, Q. Yu, Q. Wang, Z. Yuan, Hydrodeoxygenation of
[32] D. Sun, T. Kiyobayashi, H.T. Takeshita, N. Kuriyama, C.M. Jensen, X-ray diffraction
guaiacol as lignin model compound for alkanes preparation with palladium-
studies of titanium and zirconium doped NaAlH 4: elucidation of doping induced
carbon catalysts, Trans. Chin. Soc. Agric. Eng. 28 (21) (2012) 193–199.
structural changes and their relationship to enhanced hydrogen storage properties,
[7] X. Zhang, T. Wang, L. Ma, Q. Zhang, Y. Yu, Q. Liu, Characterization and catalytic prop-
J. Alloys Compd. 337 (1) (2002) L8–L11.
erties of Ni and NiCu catalysts supported on ZrO2–SiO2 for guaiacol
[33] B.H. Park, Y.-I. Kim, K.H. Kim, Effect of silicon addition on microstructure and me-
hydrodeoxygenation, Catal. Commun. 33 (2013) 15–19.
chanical property of titanium nitride film prepared by plasma-assisted chemical
[8] J. Sun, A.M. Karim, H. Zhang, L. Kovarik, X.S. Li, A.J. Hensley, J.-S. McEwen, Y. Wang,
vapor deposition, Thin Solid Films 348 (1) (1999) 210–214.
Carbon-supported bimetallic Pd–Fe catalysts for vapor-phase hydrodeoxygenation
[34] K. Lin, P.P. Pescarmona, H. Vandepitte, D. Liang, G. Van Tendeloo, P.A. Jacobs, Synthe-
of guaiacol, J. Catal. 306 (2013) 47–57.
sis and catalytic activity of Ti-MCM-41 nanoparticles with highly active titanium
[9] D.D. Laskar, M.P. Tucker, X. Chen, G.L. Helms, B. Yang, Noble-metal catalyzed
sites, J. Catal. 254 (1) (2008) 64–70.
hydrodeoxygenation of biomass-derived lignin to aromatic hydrocarbons, Green
[35] P. Sudarsanam, A. Rangaswamy, B.M. Reddy, An efficient noble metal-free Ce–Sm/
Chem. 16 (2) (2014) 897–910.
SiO2 nano-oxide catalyst for oxidation of benzylamines under ecofriendly condi-
[10] A.A. Dwiatmoko, L. Zhou, I. Kim, J.-W. Choi, D.J. Suh, J.-M. Ha, Hydrodeoxygenation of
tions, RSC Adv. 4 (86) (2014) 46378–46382.
lignin-derived monomers and lignocellulose pyrolysis oil on the carbon-supported
[36] B.G. Rao, P. Sudarsanam, A. Rangaswamy, B.M. Reddy, Highly efficient CeO2–MoO3/
Ru catalysts, Catal. Today.
SiO2 catalyst for solvent-free oxidative coupling of benzylamines into N-
[11] M. Hellinger, H.W.P. Carvalho, S. Baier, D. Wang, W. Kleist, J.-D. Grunwaldt, Catalytic
Benzylbenzaldimines with O2 as the oxidant, Catal. Lett. 145 (7) (2015) 1436–1445.
hydrodeoxygenation of guaiacol over platinum supported on metal oxides and zeo-
[37] B. Mallesham, P. Sudarsanam, B.V.S. Reddy, B.M. Reddy, Development of cerium pro-
lites, Appl. Catal. A Gen. 490 (2015) 181–192.
moted copper–magnesium catalysts for biomass valorization: selective
[12] Y.-K. Hong, D.-W. Lee, H.-J. Eom, K.-Y. Lee, The catalytic activity of Pd/WOx/γ-Al2O3
hydrogenolysis of bioglycerol, Appl. Catal. B Environ. 181 (2016) 47–57.
for hydrodeoxygenation of guaiacol, Appl. Catal. B Environ. 150–151 (2014)
[38] M.M. Ambursa, T.H. Ali, H.V. Lee, P. Sudarsanam, S.K. Bhargava, S.B.A. Hamid,
438–445.
Hydrodeoxygenation of dibenzofuran to bicyclic hydrocarbons using bimetallic
[13] M. Ishikawa, M. Tamura, Y. Nakagawa, K. Tomishige, Demethoxylation of guaiacol
Cu–Ni catalysts supported on metal oxides, Fuel (2016).
and methoxybenzenes over carbon-supported Ru–Mn catalyst, Appl. Catal. B Envi-
[39] M.H. Amin, S. Putla, S.B.A. Hamid, S.K. Bhargava, Understanding the role of lantha-
ron. 182 (2016) 193–203.
nide promoters on the structure–activity of nanosized Ni/γ-Al2O3 catalysts in car-
[14] C.R. Lee, J.S. Yoon, Y.-W. Suh, J.-W. Choi, J.-M. Ha, D.J. Suh, Y.-K. Park, Catalytic roles
bon dioxide reforming of methane, Appl. Catal. A Gen. 492 (2015) 160–168.
of metals and supports on hydrodeoxygenation of lignin monomer guaiacol, Catal.
[40] J. Marrero-Jerez, A. Murugan, I. Metcalfe, P. Núñez, TPR–TPD–TPO studies on CGO/
Commun. 17 (2012) 54–58.
NiO and CGO/CuO ceramics obtained from freeze-dried precursors, Ceram. Int. 40
[15] J. Lu, A. Heyden, Theoretical investigation of the reaction mechanism of the
(9) (2014) 15175–15182.
hydrodeoxygenation of guaiacol over a Ru(0001) model surface, J. Catal. 321
[41] C. Luo, D. Li, W. Wu, Y. Zhang, C. Pan, Preparation of porous micro–nano-structure
(2015) 39–50.
NiO/ZnO heterojunction and its photocatalytic property, RSC Adv. 4 (6) (2014)
[16] A.J.R. Hensley, Y. Wang, J.-S. McEwen, Adsorption of guaiacol on Fe (110) and Pd
3090–3095.
(111) from first principles, Surf. Sci.
[42] A.V. Bridgwater, Review of fast pyrolysis of biomass and product upgrading, Biomass
[17] H.W. Lee, B.R. Jun, H. Kim, D.H. Kim, J.-K. Jeon, S.H. Park, C.H. Ko, T.-W. Kim, Y.-K.
Bioenergy 38 (2012) 68–94.
Park, Catalytic hydrodeoxygenation of 2-methoxy phenol and dibenzofuran over
[43] M. Zhou, H. Chai, D. Jia, W. Zhou, The glucose-assisted synthesis of a graphene
Pt/mesoporous zeolites, Energy 81 (2015) 33–40.
nanosheet–NiO composite for high-performance supercapacitors, New J. Chem. 38
[18] K. Leiva, N. Martinez, C. Sepulveda, R. García, C.A. Jiménez, D. Laurenti, M. Vrinat, C.
(6) (2014) 2320–2326.
Geantet, J.L.G. Fierro, I.T. Ghampson, N. Escalona, Hydrodeoxygenation of 2-
[44] R.O.D. Fonseca, A.A.A.D. Silva, M.R.M. Signorelli, R.C. Rabelo-Neto, F.B. Noronha,
methoxyphenol over different Re active phases supported on SiO2 catalysts, Appl.
R.C.C. Simões, L.V. Mattos, Nickel/doped ceria solid oxide fuel cell anodes for dry
Catal. A Gen. 490 (2015) 71–79.
reforming of methane, J. Braz. Chem. Soc. (2014).
[19] W. Mu, H. Ben, X. Du, X. Zhang, F. Hu, W. Liu, A.J. Ragauskas, Y. Deng, Noble metal
[45] B.M. Reddy, K.N. Rao, P. Bharali, Copper promoted cobalt and nickel catalysts sup-
catalyzed aqueous phase hydrogenation and hydrodeoxygenation of lignin-
ported on ceria–alumina mixed oxide: structural characterization and CO oxidation
derived pyrolysis oil and related model compounds, Bioresour. Technol. 173
activity, Ind. Eng. Chem. Res. 48 (18) (2009) 8478–8486.
(2014) 6–10.
[46] P. Sudarsanam, B. Hillary, B. Mallesham, B.G. Rao, M.H. Amin, A. Nafady, A.M.
[20] G. Yao, G. Wu, W. Dai, N. Guan, L. Li, Hydrodeoxygenation of lignin-derived phenolic
Alsalme, B.M. Reddy, S.K. Bhargava, Designing CuOx nanoparticle-decorated CeO2
compounds over bi-functional Ru/H-Beta under mild conditions, Fuel 150 (2015)
nanocubes for catalytic soot oxidation: role of the nanointerface in the catalytic per-
175–183.
formance of heterostructured nanomaterials, Langmuir 32 (9) (2016) 2208–2215.
[21] M. Grilc, B. Likozar, J. Levec, Hydrodeoxygenation and hydrocracking of solvolysed
[47] R. Ryoo, S. Jun, Improvement of hydrothermal stability of MCM-41 using salt effects
lignocellulosic biomass by oxide, reduced and sulphide form of NiMo, Ni, Mo and
during the crystallization process, J. Phys. Chem. B 101 (3) (1997) 317–320.
Pd catalysts, Appl. Catal. B Environ. 150 (2014) 275–287.
[48] H. Shafaghat, P.S. Rezaei, W.M.A.W. Daud, Catalytic hydrogenation of phenol, cresol
[22] M. Grilc, B. Likozar, J. Levec, Hydrotreatment of solvolytically liquefied lignocellulos-
and guaiacol over physically mixed catalysts of Pd/C and zeolite solid acids, RSC Adv.
ic biomass over NiMo/Al2O3 catalyst: reaction mechanism, hydrodeoxygenation ki-
5 (43) (2015) 33990–33998.
netics and mass transfer model based on FTIR, Biomass Bioenergy 63 (2014)
[49] J. Kong, B. Li, C. Zhao, Tuning Ni nanoparticles and the acid sites of silica-alumina for
300–312.
liquefaction and hydrodeoxygenation of lignin to cyclic alkanes, RSC Adv. 6 (76)
[23] M. Grilc, B. Likozar, J. Levec, Simultaneous liquefaction and hydrodeoxygenation of
(2016) 71940–71951.
lignocellulosic biomass over NiMo/Al2O3, Pd/Al2O3, and zeolite Y catalysts in hy-
[50] H. Lee, H. Kim, M.J. Yu, C.H. Ko, J.-K. Jeon, J. Jae, S.H. Park, S.-C. Jung, Y.-K. Park, Cat-
drogen donor solvents, ChemCatChem 8 (1) (2016) 180–191.
alytic Hydrodeoxygenation of Bio-oil Model Compounds over Pt/HY Catalyst, Scien-
[24] T. Nimmanwudipong, R.C. Runnebaum, D.E. Block, B.C. Gates, Catalytic reactions of
tific Reports 6, 2016.
guaiacol: reaction network and evidence of oxygen removal in reactions with hy-
drogen, Catal. Lett. 141 (6) (2011) 779–783.
[25] E. Furimsky, J. Mikhlin, D. Jones, T. Adley, H. Baikowitz, On the mechanism of
hydrodeoxygenation of ortho substituted phenols, Can. J. Chem. Eng. 64 (6)
(1986) 982–985.

You might also like