Você está na página 1de 18

Research Article

pubs.acs.org/acscatalysis

Screening of Catalysts for Hydrodeoxygenation of


Phenol as a Model Compound for Bio-oil

Peter M. Mortensen, Jan-Dierk Grunwaldt, Peter A. Jensen, and Anker D.

Jensen*,

Department of Chemical and Biochemical Engineering, Technical University of Denmark, Sltofts
Plads, Building 229, DK-2800 Lyngby, Denmark

Institute for Chemical Technology and Polymer Chemistry, Karlsruhe Institute of Technology
(KIT), Engesserstrasse 20, D-79131 Karlsruhe, Germany
*S Supporting Information

ABSTRACT: Four groups of catalysts have been tested for hydro-


deoxygenation (HDO) of phenol as a model compound of bio-oil, including oxide catalysts, methanol synthesis catalysts, redu
investigated and ZrO2 was found to perform best. Pt/C, Ni/CeO2, and Ni/CeO2-ZrO2 were the most active catalysts for the initia

1. INTRODUCTION CH1.4O0.4 + 0.7H2 CH2 + (1


A
is prospective route for production of biofuels
the conversion 0.4H2O )

of biomass into bio-oil through ash pyrolysis that is separated from the water and ultimately
followed by upgrading via hydrodeoxygenation gives a product equivalent to crude oil. The
1
(HDO). Flash pyrolysis is reaction can be generally written as (normalized to
1
advantageous in making a locally produced feed carbon)
liquid that minimizes 24
transportation costs to
larger bioreneries.
However, bio-oil is a viscous, polar, and acidic
liquid with a low heating value, making it, in
most cases, unsuitable as an engine fuel
directly. These unfavorable characteristics are
all
associated with high levels of water (1030 wt
%) and oxygen- containing organic
5 compounds
(3040 wt % oxygen) in the oil.
In HDO, bio-oil is treated with hydrogen at a
pressure of up to 200 bar and temperatures in
the range from 200 to 400 C. This converts
the oxy compounds to a hydrocarbon product
where CH2 represents an unspecied and low pressures (<100 bar). In particular,
hydrocarbon as a product. low temperatures are desirable to prevent
One of the major challenges with this coking,
1,7
and therefore, focus has in the
concept is to nd a catalyst with a high current screening study been on catalysts
activity for the deoxygenation reaction operating at 275 C. Four categories of
and at the same time obtain a sucient catalysts were tested as catalysts for HDO of
lifetime, as deposition of carbonaceous
species has proven to be a severe Received: April 9, 2013
1,6
problem. Preferably, HDO catalysts Revised: June 26, 2013
should be relatively inexpensive and Published: July 1, 2013
function at low temperatures (<300 C)

2013 American Chemical Society 1774 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,


17741785
ACS Research
Catalysis Article

1,10 11,12
Figure 1. Proposed reaction mechanisms of (a) oxide catalysts and (b) reduced metal catalysts.

Table 1. Summary of Catalyst and Support Properties, i.e., the Theoretical Loading of the Active
Component, the Particle Size of the Support (Sieve Fraction), the Specic Surface Area, and
Reduction Conditions for the Tested Catalysts
catalyst/support loading (wt %) particle size surface reduction T (C)/t
(m) area (m2/g) (h)
carbon 1100
SiO2 250
ZrO2 160
CeO2 140
CeO2-ZrO2 140
MgAl2O4 90
Al2O3 150
MnO/C 15 7101400 680 none
WO3/C 15 7101400 790 none
MoO3/C 15 7101400 660 none
V2O5/C a
15 7101400 310 none
NiO-MoO3/Al2O3 300 none
a
CoO-MoO3/Al2O3 60
300 none
Cu/ZnO/Al2O3 45% Cu/5% Zn 60
900 60 300
NiCu/SiO2 10% Ni/10% 900 180 /7
35
Cu/SiO2 Cu
15 300 170 0/1
38
Ru/C 5 60
15 1160 0/1
400
Pd/C 5 15 1100 /2
400
Pt/C 5 15 1140 /2
400
Co/SiO2 5 300 210 /2
550
Fe/SiO2 5 60
300 200 /2
550
Ni/SiO2 5 60
631 210 /2
400
Ni/Al2O3 5 25
631 140 /2
550
Ni/CeO2 5 25
631 130 /2
400
Ni/ZrO2 5 25
631 130 /2
550
Ni/CeO2-ZrO2 5 25
631 120 /2
550
Ni/MgAl2O4 5 25
631 80 /2
550
Ni/C 5 25
631 1020 /2
550
Ni-V2O5/SiO2 5% Ni/5% V 25
300 210 /2
550
Ni-V2O5/ZrO2 5% Ni/5% V 60
300 120 /2
550
a
Commercial catalyst by Haldor Topse A/S without detailed characterization data. /2
Oxide catalysts have been proposed to catalyze
phenol: (1) oxide catalysts, (2) methanol the reaction through roughly three steps as
synthesis catalysts, shown in Figure 1a: chemisorption via the oxygen
(3) reduced noble metal catalysts, and (4) atom on a coordinatively unsaturated metal site,
reduced non-noble metal catalysts. Phenol was donation of a proton from a hydroxyl group, and
1,10
chosen as a model compound of bio-oil, as desorption. In this mechanism, the generation of
phenols have been identied among the most
8 9
persistent, yet fairly abundant, compounds in
bio-oil.
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
75
vacancyACSsites (the dotted circle in the Research oxy
catalysts could potentially activate
Article
gure) Catalysis
is responsible for the activation of compounds through
both the oxy compound and hydrogen on their oxygen group and in this way permit reaction.
1,10
the catalytic surface. Figure 1b shows the key concepts of the
Methanol synthesis catalysts have reaction mechanism for reduced metal
1
proven to be able to activate CO/CO 2 catalysts. The reaction is initiated by
1316
through the oxygen. Thus, these adsorption of the oxy compound on the
catalyst surface.
Adsorbed hydrogen on the active metal dried at 110 C for at least 12 h and then calcined
clusters reacts with the oxy compound to at 400 C with a heating rate of 10 C/min and a
facilitate the deoxygenation. This is followed holding time of 4 h. The catalysts with carbon
by desorption of the nal product. The oxy supports were calcined in nitrogen, while other
compound adsorption step can take place catalysts were calcined in air.
either on the support or directly on the active For example, NiCu/SiO2 was prepared by
metal, depending on which type of metal is dissolving 3.80 g of Cu(NO3)23H2O and 4.95 g of
used. Specically, noble metals have been Ni(NO3)26H2O in 8 mL of deionized water and
shown to be capable of activating the oxy impregnation of this solution on 8 g of dried SiO2.
compound on the metal sites.
12,17
For non- The Ni-V2O5 catalysts were made by initially
noble metals, the activation is thought to occur impregnating the support with vanadium, drying
the catalyst, and then
through an oxygen vacancy site in the support
11,18 impregnating it with nickel nitrate. NH4VO3 was
metal oxide, similar to the activation step dissolved by usinginadditionally oxalic acid (Sigma-
Aldrich, 99.0%) a
for the oxide catalysts shown in Figure 1a. molar ratio of 1:2 (V:oxalic acid).
Hence, these four dierent classes of catalysts Cu/ZnO/Al2O3 was prepared according to the
19
were screened in this work. method of Baltes et al. An aqueous solution of
Cu(NO3)23H2O (0.6 mol/L, Sigma-Aldrich, 99%),
Zn(NO3)26H2O (0.3 mol/L, Sigma-Aldrich, 99%),
2. EXPERIMENTAL SECTION and Al(NO3)39H2O (0.1 mol/L, Sigma-Aldrich,
98%) was coprecipitated with a solution of
Na2CO3 (1 mol/L, Sigma-Aldrich, 99%) for 1 h.
2.1. Catalyst Synthesis. MnO/C, WO3/C, During the
MoO3/C, V2O5/C, Cu/SiO2, NiCu/SiO2, Co/SiO2,
Fe/SiO2, Ni-based catalysts, and Ni-V-based
catalysts were all prepared by incipient
wetness impregnation of the corresponding
supports. An overview of the applied catalysts
is given in Table 1. The precursors for the
active materials were Mn(C2H3O2)24H2O
(Sigma-Aldrich, 99%), (NH4)6H2W12O40xH2O
(Sigma-Al- drich, 99.0%),
(NH4)6Mo7O244H2O (Sigma-Aldrich,
99.0%), NH4VO3 (Sigma-Aldrich, 99.0%),
Cu- (NO3)23H2O (Sigma-Aldrich, 99.0%),
Fe(NO3)39H2O
(Sigma-Aldrich, 98.0%), Fe(NO3)26H2O
(Sigma-Aldrich,
98.0%), Ni(C2H3O)24H2O (Sigma-Aldrich,
98.0%), and Ni(NO3)26H2O (Sigma-Aldrich,
97.0%).
The active carbon was Daihope 009. The
silica was supplied by Saint-Gobain NorPro
(type SS6*138 with a purity of
99.5%). Also, the ZrO2 was supplied by Saint-
Gobain NorPro
(type SZ6*152 with an impurity of 3.3% SiO 2).
The alumina was supplied by Sasol (type
Puralox TH 100/150). The spinel
was produced from the Al 2O3 by mixing
stoichiometric amounts of Al2O3 and MgO and
calcination of the mixture at 900 C; the
product was conrmed to be a mixed oxide
(MgAl2O4) by X-ray diraction (XRD). CeO2 and
CeO2-ZrO2 were supplied by AMR Ltd. CeO2-
ZrO2 was conrmed by XRD to be a mixed
oxide. Before impregnation, all supports were
crushed and sieved.
Incipient wetness impregnation was made by
initially dissolving the corresponding amount
of precursors in deionized water equivalent to
the pore volume of the support and then
mixing with the support. All precursors
described above were suciently water-
soluble. After impregnation, the samples were
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
76
ACS
precipitation process, the pH was therefore dicult with which to workResearchin the
Catalysis Article
maintained at 7 0.1 and the product separation and gas chromatography
temperature was kept at 60 C. Afterward, (GC) analysis.
the precipitate was ltered and washed In a blank experiment without catalyst, 10 g
with demineralized water followed by of phenol and 40 mL of H2O were allowed to
drying overnight at 80 C and calcining at react for 4 h at 275 C and 100 bar. Almost no
300 C under air for 3 h. catalytic activity was seen in this case as a
NiO-MoO3/Al2O3 and CoO-MoO3/Al2O3 conversion of 0.3% was observed. Thus, the
catalysts were obtained from Haldor reactor was hardly catalytically active and did
Topse A/S. Ru/C, Pd/C, and Pt/C were not inuence the experiments. The blank
obtained from Sigma-Aldrich. experiment was repeated occasionally to
Table 1 also summarizes the theoretical ensure that the reactor was not contaminated
loadings, particle sizes, specic surface over time.
areas, and pretreatment conditions of the By 3-fold repetition of a hydrodeoxygenation
dierent catalysts. experiment with Ni/SiO2, it was found that this
2.2. Catalyst Testing. The screening experiments procedure typically had an uncertainty in the
were measured yields of 2 mol %, corresponding
performed in a 300 mL batch reactor (Parr, type to <5% as the relative standard deviation.
4566) made Overall, the repeatability of the experiments
from Hastelloy C steel. One gram of was good.
catalyst was loaded in the reactor, and then In some cases, shorter experiments had to
50 g of phenol (Sigma-Aldrich, 99%) was be performed not to reach 100% conversion.
added. The mixture was stirred with a
propeller at 380390 rpm throughout the Here the batch reactor was initially heated
experiment and heated to 275 C in a without its contents being stirred. It was
hydrogen atmosphere, giving a nal
pressure of 100 bar. The heating rate was assumed that the extent of the reaction was
12.5 C/min. During the experiments, low in the heating phase because of the mass
hydrogen was added continuously to transfer restriction of hydrogen in such
maintain the pressure. To 20
stop the experiment, the reactor was systems.
placed in an ice bath (cooling rate of This was supported by the observation that no
2550 C/min). The start of the hydrogen was consumed in the heating phase.
experiment was taken as the time when
the heater was turned on and the When the desired temperature
end of the experiment when the reactor was reached, the stirring was started at 380390 rpm
was lowered into the ice bath. The and the
resultant product was ltered by suction experiment could be performed at close to
ltration to separate the product liquid isothermal
and catalyst. conditions. This made it possible to measure
Catalytic activity measurements over the activity in short-term (530 min)
experiments.
oxide and methanol synthesis catalysts As specied in Table 1, some of the catalysts were
were performed with a mixture of 10 g of pretreated
phenol in 40 mL of deionized water as in hydrogen to reduce the active metals. This
feedstock, because low levels of was done in a continuous ow setup, where
conversion were observed in this series of the sample was treated at the specied
experiments. Addition of water was temperature in a 50:50 mixture of hydrogen
required as a solvent for phenol, as the and nitrogen at a total ow of 500 NmL/min.
reactant is solid at room temperature and Temperature- programmed reduction by
hydrogen (H2-TPR) was used to
evaluate the required reduction temperature,
and XRD conrmed that the catalysts were The analysis was performed on the basis of
reduced under the specied conditions. changes in mass, and the total mass applied
was on the order of 2040 mg. Five volume
2.3. Product Analysis. Analysis of the percent hydrogen in nitrogen was added at a
liquid product was performed with a ow of 100 NmL/min, and the samples were
Shimadzu GCMS/FID-QP2010UltraEi heated at a rate of 5 C/ min to 300 C, at a
rate of 1 C/min from 300 to 500 C, and
instrument tted with a Supelco Equity-5 at a rate of 5 C/min from 500 to 700 C.
column and equipped with a mass
spectrometer (MS) for product identication NH3-TPD was performed with an Autochem II
2920 apparatus. Here 0.1 g portions of the
and a ame ionization detector (FID) for samples were initially
heated to 500 C in a 50 mL/min He ow and
quantication. External standards were then cooled to 100 C. At this point, the
prepared for phenol, cyclohexanol, cyclo- samples were saturated with a 50 mL/min NH3
hexanone, and cyclohexane using ethanol as a ow for 120 min. Desorption of NH3 was
solvent. The concentrations of the remaining hereafter measured by ushing with a 50
peaks were calculated from the FID on the mL/min He ow while heating the sample at a
basis of the eective carbon number method,
21 rate of 5 C/min to 500 C.
X-ray diraction (XRD) was measured with a
where the concentration of a compound is PANalytical
found to be
C =C Ai e,ref
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
77
ACS
XPe PRO Research
ument using a rotating copper anode X-ray
rt
Catalysis instr Article
re
i ref
Af e, (2) source at 40 kV and 30 mA, a nickel lter, and
i
automatic

where C is the concentration, A the area of the antiscatter and divergence slits. The 2 angle
peak in the FID spectrum, and e the eective was scanned from 12 to 120 in increments of
carbon number. Index i refers to the compound 0.00656 with 45.9 s per step. The mean
with the unknown concentration, and index ref coherent-scattering domain (CSD) size was
refers to a reference compound where the derived from
concentration is known. In all calculations reection half-widths using the SelyakovScherrer
with this formula, cyclohexanol was used as a 23
reference. The eective carbon number was formula.
obtained from the review of Schoeld.
21 All samples were analyzed as prepared powders.
The conversion, X, was calculated as
3. KINETICS OF PHENOL HDO
n phenol To determine the activity of the catalysts on a
X = 1 100%
quantitative basis, a kinetic model was
developed. In the following, this model is
presented as a basis for the later discussion.
The time-dependent development of the
conversion of
(3) phenol and the yields of cyclohexanone,
cyclohexanol, and cyclohexane in an
n0,phenol experiment with a Ni-V2O5/SiO2 catalyst at
where nphenol is the moles of phenol after 250 C and 100 bar is shown in Figure 2. At rst, phenol
reaction and n0,phenol was
the moles of phenol prior to reaction.
The yields (Yi) of relevant products were calculated as
ini
Yi = 100%
6 0,phenol (4)
n
where ni is the moles of product i after reaction
and i the number of carbon atoms in compound
i.
The carbon balance was evaluated in all
experiments by comparing the carbon initially
in the reactor to the carbon measured in the
product and on the catalyst:
n 6n
i i i 0, phenol
C = 6n 100%
0,phenol (5)

where C is the carbon deviation in percent Figure 2. Conversion of phenol (X) and yields of
and ni the moles of compound i. All compounds cyclohexanone, cyclohexanol, and cyclohexane as a
function of time over a Ni-V2O5/ SiO2 catalyst. The
identied in the GC analysis were included in experiments were conducted with 1 g of catalyst in
the carbon balance. Generally, the carbon 50 g of phenol. T = 250 C, and P = 100 bar.
balance was closed within 10%, but in
many experiments, almost

complete closure was achieved; this will be discussed later.

All calculations of X, Yi, and C were corrected


for loss of mass from transferring and ltration
processes, which was
quantied as the average of a number of blank Temperature-programmed reduction by
tests. hydrogen was performed for selected catalysts on
2.4. Catalyst Characterization. A a Netzsch STA 449 F1 Jupiter ASC
Quantachrome iQ2 instrument was used for thermogravimetric analyzer to identify the
measurement of the specic surface area on required temperature for complete reduction of
22
the basis of BET theory. Nitrogen at its the supported metal.
boiling point was used in the p/p0 range from
0.05 to 0.3 to construct a seven-point BET
plot. Generally, all the tested supports had
surface areas on the same order of magnitude;
only the carbon support had a signicantly
larger surface area (see Table 1).
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
78
ACSto primarily cyclohexanol. This
converted deoxygenation, with cyclohexanol Research
as the
Catalysis Article
yield was maximal after reaction for intermediate product and cyclohexane as the
approximately 3.5 h. The cyclohexane nal product. Also, cyclohexanone and
yield, on the other hand, increased cyclohexene were observed in small amounts.
continuously throughout the experiment, The cyclohexanone yield reached a maximum
with an increasing rate toward the end. of 1.7% after 1 h and then steadily decreased
The data in Figure 1 indicate a reaction mechanism (cf. Figure 2). The cyclohexene yield was even
at 250 less and on the order of 0.01%.
C in which hydrogenation of the aromatic
ring occurs as a rst step followed by
24
Shuikin and Erivanskaya suggested that aomatic alcohol (469 kJ/mol)
hydrogenation of phenol over Ni catalysts
produces cyclohexanone as the primary secondary alcohol (385 kJ/mol)
product; however, the subsequent
primary alcohol (383 kJ/mol)
hydrogenation of cyclohexanone to
cyclohexanol proceeds at a much higher rate, tertiary alcohol (379 kJ/mol) (6)
and therefore, cyclohexanone is seen in only
low yields. Similarly, cyclohexene is formed The data clearly show that the aromatic
through dehydration of cyclohexanol followed alcohol (i.e., phenol) has an 100 kJ/mol
by quick hydrogenation to cyclohexane. higher bond dissociation energy than the
secondary alcohol (as cyclohexane), and the
Overall, the observations from Figure 2 and deoxygenation can
24
previous literature for nickel catalysts, therefore more readily take place from the
25,26 27 28
Pd/C, Pt/C, Ni/HZSM-5, and cyclohexanol compared to the phenol as the
29 CO bond is markedly weakened.
Ni-MoS2/Al2O3 suggest the reaction scheme Because the rates of hydrogenation of both
depicted in Figure 3 with solid arrows. To cyclohexanone and cyclohexene are relatively
verify this, measurements were
fast and the yields of these compounds are
low, the eective reaction scheme can be
described by the dotted arrows in Figure 3,
which results in the following two overall main
reactions:
phenol + 3H2 cyclohexanol (I)

Figure 3. Observed reaction path for HDO of cyclohexanol + H2 (II


phenol. Solid arrows indicate main pathways, cyclohexane + H2O )
while dashed arrows show the eective
conversion steps in the kinetic model. Indicated Assuming rst-order dependency of the
rates are from measurements with a 5 wt % Ni/ZrO 2 hydrocarbon com- pounds leads to the
catalyst at 275 C and 100 bar in the batch reactor
setup. The estimations of the individual turnover following rate expressions:
n
frequencies (TOF) are supplied in the Supporting r1 = k1Cphenol
H
P (7
Information. 2 )
r2 = k2CC
H
m
hexanolP
2
(8
performed over a Ni/ZrO2 catalyst to quantify )
the rate of the
four steps (see the Supporting Information). where ri is the rate of reaction i, ki is the rate
This is shown in constant of
Figure 3 as the turnover frequency (TOF) in reaction i, Ci is the concentration of either
each step. Hydrogenation of cyclohexanone phenol or cyclohexanol, and n and m are the
and hydrogenation of cyclo- hexene were reaction orders of hydrogen in reactions I and
signicantly faster reactions than II, respectively. The assumption of rst-order
hydrogenation of phenol and dehydration of kinetics with respect to the hydrocarbons is
cyclohexanol. Hence, only low yields of made in the absence of a better estimate. As
cyclohexanone and cyclohexene should be all the experiments were performed with the
expected. same and constant pressure of hydrogen,
From the data in Figures 2 and 3, the hydrogen pressure can be included in the
deoxygenation appears to take place more rate constant and thereby give the following
readily from a saturated ring than from an lumped rate expressions:
unsaturated ring, which has also been shown
elsewhere.
1,8,30 r1 = k1Cphenol (9)
This is related to the dissociation
31 energy of the
CO bond in alcohols, which decreases in
the following order:

17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,


79
ACS Research
Catalysis Article

Figure 4. Conversion of phenol (X) and yields of cyclohexanone, cyclohexanol, and cyclohexane from
experiments with dierent types of oxide catalysts (Oxide) and methanol synthesis catalysts (Methanol
Synthesis) and for the blank experiment. The experiments were conducted with 0.5 g of catalyst in 10 g of
phenol and 40 mL of water. T = 275 C, and P = 100 bar. The reaction time was 4 h.
r2 = k 2CChexanol
ZnO/Al2O3). This group of catalysts also had a low
activity.
(10) Cu/ZnO/Al2O3 had a conversion of 9% but
primarily was selective toward hydrogenation,
where k1 is the rate constant for the with cyclohexanol and cyclo-
hydrogenation reaction and k2 the rate hexanone constituting 91% of the product.
constant for the deoxygenation reaction. The NiCu/SiO2 had a conversion of 7%, but
rate expressions in eqs 9 and 10 were
combined with the batch reactor design according to the carbon balance (cf. Table 2),
32
equation to allow determination of k1 and k2 half of the measured conversion was not
from the experimental data and a quantitative accounted for.
comparison of the dierent catalysts. Among these catalysts, Cu/ZnO/Al 2O3 was
A thorough description of the derivation of the best performing, with the highest
the model, the assumptions made, and the conversion and a good closure of the carbon
validation of the model can be found in the balance (C = 0.2%). However, as the
product primarily was cyclohexanol, the
Supporting Information. potential of this catalyst seems
limited in the context of HDO.
4. RESULTS AND DISCUSSION Ardiyanti et al.
34
previously investigated
4.1. Oxide and Methanol Synthesis Catalysts. Cu/Al2O3 and NiCu/Al2O3 catalysts at varying
The results from testing the oxide catalysts at Ni/Cu ratios for HDO of anisole at 300 C and
10 bar in a continuous ow reactor. In
275 C and 100 bar are shown in Figure 4 in these studies, it was observed that the pure Cu
terms of the yields of cyclohexanone, catalyst was unable to perform HDO of the
cyclohexanol, and cyclohexane, and the anisole, as the only product was phenol. For
conversion of phenol. None of the catalysts NiCu/Al2O3, the best catalyst was obtained
achieved a conversion of >10%. The NiO- with a 8:1 Ni:Cu weight ratio. In agreement
MoO3/Al2O3 catalyst showed some with this, we also observed the improved
hydrogenation activity with a yield of 4% activity of the NiCu/SiO2 (with a Ni:Cu ratio of
oxygenated cyclohexanes. However, the 1) catalyst relative to that of Cu/SiO2, but the
activity for HDO over this catalyst was low, activity of NiCu/SiO2 was not optimized as in
with a yield of cyclohexane of only 0.9%. the work of Ardiyanti et al.
34
The results
Table 2 gives a summary of the performance indicate that the Ni loading has to be markedly
of the catalysts. The carbon balances of the higher than the Cu loading to achieve good
experiments were reasonably well HDO activity.
In summary, the apparent order of activity
Table 2. Overview of the Results from for the tested oxide and methanol catalysts
Dierent Oxide Catalysts and Methanol was as follows:
a
Synthesis Catalysts
k1 (mL kg k2 (mL ca
kg 1 Cu/ZnO/Al2O3 NiOMoO3/Al2O3
catalyst
1
ca
min1t ) C NiCu/SiO2
min1t) (%)

blank
> Cu/SiO2 > MnO/C, WO3/C, V2O5/C
0.0
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
80
ACS
MnO/C 1 7 1.1 Research
Catalysis
WO3/C 1 10 1.4 , CoOMoO3/Al2O3 Article
MoO3/C 2 4 3.1
V2O5/C 1 3 1.4 Commonly, none of the catalysts achieved high
NiO-MoO3/Al2O3 5 20 2.2 conversion under these conditions because of
CoO-MoO3/Al2O3 6 0 8.5 their inability to hydrogenate the aromatic
Cu/ZnO/Al2O3 8 10 0.2 ring of phenol (cf. Table 2).
NiCu/SiO2 7 28 3.9 4.2. Reduced Noble Metal Catalysts. In the
Cu/SiO2 1 50 0.1 experiments with reduced noble metal
a
k1 is the rate constant for hydrogenation. k2 is the catalysts, higher conversions were observed
rate constant for deoxygenation. C is the deviation compared to the oxide and methanol synthesis
in the carbon balance. The experiments were catalysts. Thus, pure phenol was used for
conducted with 0.5 g of catalyst in 10 g of phenol these experiments, as this gave consistency.
and 40 mL of water. T = 275 C, and P = 100 bar.
The reaction time The reaction temperature and pressure were
was 4 h. kept the same.
Figure 5 summarizes the conversion of
phenol and the yields of cyclohexanol,
closed and in most cases were on the order of cyclohexane, and dicyclohexyl ether from the
2%. Only the CoO-MoO3/Al2O3 catalyst had a
larger carbon deviation of experiments with noble metal catalysts. All the
8.5%. As the conversion of phenol was 9% on catalysts gave complete conversion of the
this catalyst, it phenol through a relatively fast hydrogenation
could indicate that this catalyst mainly33was reaction to cyclohexanol. The subsequent HDO
active for cracking. Prasomsri et al. did more favorably took place on the ruthenium
observe HDO activity of bulk V2O5, Fe2O3, CuO, catalyst than on the platinum and palladium
WO3, and MoO3 catalysts in a continuous ow catalysts. The ruthenium catalyst provided a
reactor for acetone deoxygenation in the gas
phase at atmospheric pressure and 400 C, with yield of 52% cyclohexane relative to 1 and 11%
MoO3 performing the best. In contrast to their on the platinum and palladium catalysts,
study for which temperatures of 400 respectively. The palladium catalyst
C were required to deoxygenate anisole, furthermore had a high yield of dicyclohexyl
much lower temperatures were used in the ether of 21%, i.e., higher than the cyclohexane
current screening. The higher temperature yield.
requirement is probably linked to the inability Table 3 summarizes the kinetic parameters
of the catalyst to hydrogenate the anisole and of the reactions. The hydrogenation rate
instead being forced constants (k1) were on the same order of
to break the stronger CO bond of the magnitude for Ru/C and Pd/C, but Pt/C had a
rate of hydrogenation 1 order of magnitude
aromatic alcohol higher. In general, all three catalysts, however,
compared to the saturated form, as discussed were good hydrogenation catalysts,
in section 3 (see achieving 100% conversion of the phenol in
eq 6). the 5 h experiments.
Figure 4 also shows the results of the With regard to the hydrodeoxygenation
screening of the methanol synthesis catalysts step (k2), the catalytic activity of Ru/C was
(Cu/SiO2, NiCu/SiO2, and Cu/ 1 order of magnitude higher than that of Pd/C
and 2 orders of magnitude higher than that

17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,


81
best performing catalyst for guaiacol HDO in
comparison to platinum and palladium.
39
Adriyanti et al. found Pd/ZrO2 to be a better
catalyst for bio-oil HDO in a batch reactor
compared to Pt/ZrO2. Thus, these results for
HDO over noble metal catalysts are in good
agreement with the literature, which all point
toward ruthenium being one of the best
performing noble metals for HDO and with Pd
being more ecient than Pt. The kinetic model
revealed that the reason for this is that Pt is a
relatively poor deoxygenation catalyst but a
very good hydrogenation catalyst.
The reason for ruthenium being the best
performing noble metal could be correlated to
the anity of the metals to bind oxygen.
40
Nrskov et al. showed through density
functional theory (DFT) calculations that the
Figure 5. Conversion of phenol (X) and yields of binding energies of oxygen
cyclohexanol, cyclohexane, and dicyclohexyl ether relative to water (EO) were 0.01, 1.53, and 1.57 eV on
from experiments with dierent types of reduced Ru,
noble metal catalysts. The experiments were
conducted with 1 g of catalyst in 50 g of phenol. T Pd, and Pt, respectively. Thus, ruthenium has the
= 275 C, and P = 100 bar. The reaction time was strongest
5 h. binding energy with respect to oxygen, and
platinum the weakest, which correlates with
Table 3. Overview of the Results from the the anity for performing deoxygenation.
Reduced Noble Metal Catalysts
a In this work, the eect of support was not investigated,
but
previous work has shown that the type of
support can signicantly increase the activity
1 of noble metal catalysts. Lee et
min1) k (mL kg 1 min1) C (%)
catalyst dM (nm) k1 38
2 al. tested noble metal catalysts on C, Al2O3, and SiO2-
(mL kgcat
Al2O3
cat
b
Ru/C 7 1950 115 identied in the oil phase. This shows that cracking
10.8 reactions take place on the Ru/C catalyst. It has
c
Pt/C 4 31200 1 previously been shown that Ru/C can produce CH4
11.8 and CO2 from guaiacol at temperatures above 250
d 35 35
Pd/C 6 1840 19 C. Elliott and Hart linked the production of
6.1 these gases to aqueous phase re-forming, which for
a
dM is the metal crystallite size measured by Ru/C has been shown to occur under similar
36
XRD. k1 is the rate constant for hydrogenation. k2 conditions. Overall, the apparent order of activity
is the rate constant for deoxygenation. C is the for deoxygenation for the tested noble metal
deviation in the carbon balance. The experiments
were conducted with 1 g of catalyst in 50 g of catalysts for HDO of phenol was found
phenol. T = 275 C, and P = 100 bar. The reaction to be
b
time was 5 h. Determined from an 18 min
isothermal experiment at 275 C with 50% phenol
c
conversion. Determined from a 10 min isothermal
Ru/C > Pd/C > Pt/C
experiment d at 275 C with 99.8% phenol 35
conversion. Determined from a 16 min isothermal In agreement with this, Elliott and Hart found
experiment at 275 C with 44% phenol conversion. Ru/C to be a better catalyst than Pd/C for the
conversion of guaiacol in a batch reactor.
37
Wildschut et al. identied Ru/C and Pd/C as
better performing catalysts compared to Pt/C for
of Pt/C. This analysis does not take into 38
account the fact that the primary part of the bio-oil HDO in a batch reactor. Lee et al.
identied Ru/SiO2-Al2O3 as the
deoxygenation for the Pd/C catalyst was taking
place through the ether-forming reaction,
and k 2 is therefore not completely
descriptive of the anity for deoxygenation
on Pd/C. Nevertheless, the trend is clear.
The carbon balance in Table 3 shows that all
the tested catalysts lacked on the order of
510% carbon. Analysis of the gas phase from
the Ru/C experiment by sampling in a gas bag
and then analyzing on a GC-TCD instrument
showed the presence of CO2, CH4, and CO.
Furthermore, traces of pentane, hexane,
ethanol, methanol, and acetone could be
178 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
0
for guaiacol HDO at 250 C and 4070 bar in a batch For the nickel-based catalysts, the
reactor. theoretical mass loss is 1.4 wt % for a nickel
Their work showed that the activity of all the loading of 5 wt %, assuming that all nickel is
supported noble reduced from NiO to Ni. Generally, the
metal catalysts was improved in the order measured mass loss
of increasing support acidity, i.e., an
apparent order of activity of NM/SiO2- Table 4. Results from H2-TPR of the Reduced
Al2O3 > NM/Al2O3 > NM/C, where NM is Non-Noble Metal Catalysts
a

a noble metal.41 Similar,


Foster et al. also observed increasing catalytic s e mass theoretic
activity for catalyst t n loss al
a d (wt mass
HDO of m-cresol over platinum catalysts Co/SiO2 2 5 1 1.4
with increasing acidity of the support at Ni/SiO2 6
3 2
3 .1 1.4
260 C and 1 atm in a packed bed reactor Ni/Al2O3 0
3 6
5 .
0 1.4
setup. Both studies point toward the Ni/CeO2 1
2 0
3 .1 1.4
possibility that the use of an acidic Ni/CeO2- 8
2 3
5 .1 1.4
support can increase the activity of the ZrO
Ni/ZrO
2 2
9
3 0
4 .1 1.4
1
3 8
5 .0 1.4
noble metal catalysts. The role of support Ni/MgAl2O
Ni/C 0
3 0
5 .1 1.4
is discussed further in section 4.3.3. 4

Ni- 6
3 0
3 .3 5
4.3. Reduced Non-Noble Metal Catalysts. 4.3.1.
Reduc- tion Temperature. Prior to the V2O5/SiO2 0 9 .
a
activity tests of the non-noble metal Start and end indicate the temperature interval
catalysts, the required reduction over which reduction was observed. Conditions:
gas, 5% H2 in N2; ow rate of , 100 NmL/ min;
temperatures were determined by H2-TPR.
heating ramp, 5 C/min to 300 C, 1 C/min from
Table 4 summarizes the results of these 300 to 500
measurements, showing the temperature
C, and 5 C/min from 500 to 700 C; sample amount, 2040
intervals where the reduction took place.
mg.

178 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,


0
Figure 6. Conversion of phenol (X) and yields of cyclohexanol and cyclohexane from experiments with
dierent types of reduced non-noble metal catalysts. The experiments were conducted with 1 g of catalyst in
50 g of phenol. T = 275 C, and P = 100 bar. The reaction time was 5 h.
was close to the theoretical mass loss. The largest Table 5. Overview of the Results from Reduced Non-
deviation was seen in the case of Ni/Al2O3 and Noble Metal Catalystsa
Ni/MgAl2O4, with measured mass losses of 0.9 and 0.6
N
wt %, respectively.
extent wasForinteracting
these with the supports, H
forming a mixed oxide with the alumina or 3
spinel. These oxides have been shown to k1 (mL 1 k2 (mL adsor
retard reduction, resulting in the observed dM kgcat kg1cat C
42
lower degree of reduction. catalyst (nm) min1) min1) weak
(%) strong
For Ni-V2O5/SiO2, complete reduction of the Co/SiO 2 1 5 52 2.
2
1 1 64
vanadium into NiV/SiO2 would nickel
result and
in a Fe/SiO2 6
0.
of 5 wt %. However, only a 3.1 wt mass
theoretical % weight
loss Ni/SiO2 92 230 159 3
4. 4
showing that theloss was observed,
catalyst was not Ni/Al2O3 7 0
b
100 2
0. 0
2 2
c
completely
probably still contained reduced
vanadium in and
the Ni/CeO2 8 00 6 8.
5 1
2
4
700 C. form of oxides even at Ni/CeO2- 6 1050 1 0 02
d
Overall, the identied end temperatures ZrO2 7 0 3 6. 2 5
0 fe 36 3 1 0
analyses are judged to be sucient for Ni/MgAl 1 0 25 4. 1 1
g
in Table 1, was set 2O4
Ni/C 0
1 3 1 2. 4
3 5

higher for the individual samples to ensure Ni- 4 1160 136 0
0. 0
6
4.3.2. Reduced Non-Noblecomplete reduction.
Metals. Three V
Ni-
2 O5 /SiO 2 701 256 8
1. 0
1 2
reduced non- V2O5/ZrO2 5 3 3

noble metals were tested: Ni/SiO2, Co/SiO2, a


and Fe/SiO2. These were chosen on the basis dM is the metal crystallite size measured by
of their known hydrogenation activity in other XRD. k1 is the rate constant for hydrogenation. k2
4346 is the rate constant for deoxygenation. C is the
processes. Figure 6 shows the comparison deviation in the carbon balance. NH3 adsorption is
between the activities of these three catalysts. the ammonia saturation measured by NH3 -TPD.
Ni/SiO2 was the most active catalyst with a The data are distinguished as weak and strong
conversion of 80%, compared to 3.2 and 0.8% adsorption, with weak being the signal recorded for
desorption peaks below 200 C and strong being
for Co/SiO2 and Fe/SiO2, respectively.
above. The experiments were conducted with 1 g of
The kinetic data for the three catalysts are catalyst in 50 g of phenol. T = 275 C, and P = 100
b
listed in Table 5. The k1 rate constants of bar. The reaction time was 5 h. Determined from a
13 min isothermal experiment at 275 C with 54%
Co/SiO2 and Fe/SiO2 were 5 and 1 c
phenol conversion. Determined from a 10 min
mL kgcat1 min1, respectively, showing that isothermal experiment
d
at 275 C with 98.7% phenol
Co/SiO2 was the better hydrogenation catalyst conversion. Determined from a 15 min isothermal
of the two. However, k1 was 2 orders of experiment
e
at 275 C with 95% phenol conversion.
magnitude larger for Ni/SiO2 in comparison. k2 Determined from a 15 min isothermal experiment
was on the same order of magnitude for at 275 f
Co/SiO2 and Fe/SiO2, but signicantly larger C with 26% phenol conversion. Determined from a 10
for Ni/SiO2, as seen from Table 5. min
XRD of the spent Ni/SiO2, Co/SiO2, and Fe/SiO2
showed the active metals to be reduced and to
have similar metal crystallite sizes.
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
81
In summary, the apparent order of activity isothermal experiment at 275 C with 74%
phenol conversion.
for the three non- noble metal catalysts is g
Determined from a 4 h experiment at 275 C with
cyclohexanol as the feed instead of phenol.

atmospheric pressure and 350 C and did not


nd any notable 1
1
Ni/SiO2 Co/SiO2 > Fe/SiO2 activity. Yakovlev et al. investigated both cobalt and
nickel
catalysts for HDO of anisole at 300 C and 10
(11) bar of hydrogen and found that Co/SiO2 was
The better performance of Co/SiO2 compared practically inactive compared to Ni/SiO 2,
to that of Fe/ SiO2 is due to a better activity for which is similar to the results of this work.
hydrogenation on this catalyst. Finding similar 4.3.3. Support Eect on Nickel Catalysts.
47
results, Filley and Roth investigated cobalt With nickel having been identied as the best
and iron supported on alumina for HDO of performing of the tested non-
guaiacol at
noble metal catalysts, this was chosen for
further investigation with respect to the NH3-TPD was performed, giving information
inuence of support. about the total amount of available acid sites
Seven dierent types of supports of on the catalysts (see Table 5). The total acidity
comparable specic surface areas (cf. Table 1), of the supports decreases in the following
were tested and very dierent activities were order:
observed, as summarized in Figure 6. Ni/C Ni/Al2O3 Ni/ZrO2 Ni/CeO2 > Ni
was the only catalyst with practically no
activity and a conversion of only 2%. On the V2O5/ZrO2
other hand, Ni/CeO2, Ni/CeO2-ZrO2, and > Ni/MgAl2O4 > Ni/CeO2ZrO2 Ni
Ni/MgAl 2O4 were very active for
hydrogenation with conversions of 100%, but V2O5/SiO2
less active for deoxygenation with yields of > Ni/SiO2 > Ni/C
cyclohexane of 4, 8, and 12%, respectively.
Ni/Al2O3 had a high activity for both Via comparison of this to the apparent order of
hydrogenation and deoxygenation with a activity discussed above (or see Table 5), it
conversion of 100% and a yield of 46% follows that the amount of available acid sites
cyclohexane. Ni/ (as measured by NH3-TPD) and the HDO
SiO2 gave a yield of cyclohexane on the same activity are not directly correlated. Most
order of magnitude (38%) but was signicantly notable is Ni/SiO2, which practically contains
less active for hydro- genation with a no acid sites but is found to be one of the three
conversion of 80%. Ni/ZrO2 was the overall best HDO catalysts (of the investigated nickel
best performing catalyst under these
conditions with almost complete conversion of catalysts). However, comparing only the rate
phenol (X = 99.8%) and a high selectivity constant for hydrogenation to the acidity gives
toward cyclohexane (YC-hexane = 83%). some consistency as it is seen that the poor
The kinetic data in Table 5 show that hydrogenation catalysts (Ni/SiO2 and Ni/C)
Ni/CeO2 had the highest hydrogenation rate have very little acidity, while the catalysts
constant (k1) followed by Ni/CeO2- with better hydrogenation anity all have
ZrO2. Both catalysts showed hydrogenation
higher acidity. However, the acidity
rate constants 1 order of magnitude higher
than those of Ni/MgAl2O4, Ni/ Al2O3, and measurements do not reveal the true acidic
Ni/ZrO2, decreasing in that order. Ni/SiO2 had nature of the catalyst under experimental
a hydrogenation rate constant 2 orders of conditions, as the reducing atmosphere in the
magnitude lower than that of Ni/CeO2. experiments will result in an increased level
In contrast, the Ni/CeO2 catalyst had of formation of oxygen vacancy sites and
practically no activity for deoxygenation. The thereby acid
deoxygenation activity of Ni/CeO2- sites.
4851
A better understanding of this is
gained by plotting the hydrogenation activity
relative to the metaloxygen bond strength
[E(MO)] of the support, as shown in Figure 7.
This
ZrO2 and Ni/MgAl2O4 was 25 times larger than for Ni/CeO2,
but still low yields of cyclohexane were produced. Only Ni/
ZrO2, Ni/SiO2, and Ni/Al2O3 had relatively high
rates of deoxygenation, decreasing in that
order. An interesting aspect of this is that
Ni/SiO2 had a higher rate constant for
deoxygenation than Ni/Al2O3, but because
Ni/Al2O3 had a higher rate of hydrogenation,
17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,
82
the cyclohexane yield with the Ni/Al 2O3
catalyst was higher than that with the Ni/SiO2
catalyst.
Ni/SiO2 and Ni/ZrO2 were additionally tested
as catalysts where the support had initially
been promoted by vanadium before
impregnation with nickel. The results of these
experiments are also shown in Figure 6, while
kinetic data are listed in Table 5. For Ni-
V2O5/SiO2 the rate of hydrogenation
was signicantly increased compared to that of the pure Ni/
SiO2 catalyst. The cyclohexane yield also increased, which was
primarily caused by the increased rate of Figure 7. Hydrogenation activity as a function of the
hydrogenation, as the metaloxygen
rate of deoxygenation did not change between bond strength [E(MO)] of the pure support. Only
the Ni-V2O5/ SiO2 and Ni/SiO2 catalysts (cf. hydrogenation rate constants for nickel catalysts
supported on pure oxides from Table 5 are plotted.
Table 5). The Ni-V2O5/ZrO2 catalyst performed E(MO) data from ref 53.
slightly worse than the non-vanadium-doped
catalyst, as both the yield of cyclohexane and
the rate constants were lower. Overall, it was reveals that oxide supports with low
metaloxygen bond energy had the highest
observed that addition of vanadium primarily activity for hydrogenation. Low MO bond
inuenced the hydrogenation rate of the energies can be linked to the tendency for
catalyst. oxygen 52,53
vacancy site generation in the
oxide, and these vacancy sites will function
The carbon balances listed in Table 5 show as Lewis acid sites.
51,53
Thus, the best
an adequate performing
closure in most cases. The largest deviation hydrogenation catalyst is therefore the
was found for the Ce-containing supports, catalyst with the strongest tendency to form
where 68% of the carbon was unrecovered. Lewis acid sites in the oxide structure that is
This indicates that either cracking or most likely on the supports with the lowest
coke
deposition could take place on these catalysts. metaloxygen bond energy. Because of this eect,
On the other hand, for the two best performing introducing
catalysts, Ni/ZrO2 and Ni/ Al2O3, the carbon V2O5 to a SiO2 support causes formation of more vacancy
balance deviation was on the order of 1%, sites
indicating that these catalysts were less prone because of the lower metaloxygen bond
energy of V2O5 (12.3 eV) compared to that of
to coke deposition. In general, better closure SiO2 (34.2 eV) and thereby increases the
of the carbon balance was found for the nickel- hydrogenation activity (cf. Table 5). In
based catalysts than for the noble metal contrast, doping
catalyst (cf. Tables 3 and 5), showing that ZrO2 with V2O5 does not induce any large changes as the
these catalysts have a weaker tendency to
cause cracking.

17 dx.doi.org/10.1021/cs400266e | ACS Catal. 2013, 3,


83
Figure 8. Proposed reaction mechanism for HDO of phenol over an oxide-supported nickel catalyst (here
Ni/ZrO2). Gray spheres represent nickel atoms.

kgcat1 min1). The dierences are probably related


to supportmetal interaction. However,
metaloxygen bond energies are relatively Ni/MgAl2O4, Ni/CeO2- ZrO2, and Ni/CeO2 all have
close (14.2 and 12.3 lower k2 values. For the Ni/
eV, respectively). MgAl2O4 catalyst, TPR showed that it was hard to
A similar correlation to the metaloxygen reduce (cf. Table 4), indicating that part of the
bond energy could not be made with respect to nickel is interacting with the support, which could
the deoxygenation anity, which indicates
that this reaction does not rely on the be the reason for the poor performance of this
activation on catalyst. For the catalysts with Ce- containing
oxygen vacancies in the support.
supports, the carbon balance indicated a tendency
Taking the yield of cyclohexane as the
primary parameter for the ranking, the toward cracking/coke formation, and therefore, the
apparent order of activity for HDO of phenol poor deoxygenation performance of these catalysts
with nickel on dierent supports can be could be related
summarized as:

Ni/ZrO2 > NiV2O5/ZrO2 > NiV2O5/SiO2 >


Ni/Al2O3
> Ni/SiO2 Ni/MgAl2O4 > Ni/CeO2ZrO2

Ni/CeO2 Ni/C

It may be expected that the nickel crystallite


particle size will inuence the kinetics of the
reactions, so it is important to note that the
ranking presented here was made for catalysts
with roughly the same nickel crystallite size
(cf. Table 5).
Comparing k2 values for Ni/ZrO2, Ni-
V2O5/ZrO2, Ni-V2O5/ SiO2, Ni/Al2O3, and Ni/SiO2
shows that all the values are on the same
order of magnitude (ranging from 100 to 361
mL
to potential deactivation of the deoxygenation, but with dierent rates. In
deoxygenating sites. No further contrast, the carbon-supported catalyst is
investigations into these issues have been active for only the deoxygenation reaction. In a
performed. continuation of the discussion in the
Obviously, Ni/C exhibited behavior Introduction, it follows that the dual action of
markedly dierent from that of the other the catalyst primarily is required in the
supports. Hence, this catalyst was tested hydrogenation of complex molecules as
in an experiment in which 50 g of phenol. On the other hand, the deoxygenation
cyclohexanol was used as feed with 1 g of can take place on the nickel surfaces. Thus, we
catalyst at 100 bar and 275 C for 4 h. In suggest a mechanism as sketched in Figure 8
this experiment, a conversion of 47% was for the HDO of phenol over nickel catalysts.
observed with a yield of 42% cyclohexane, The coordinatively unsaturated metal sites in
the remaining product being primarily the oxide surface often behave like Lewis
dicyclo- hexyl ether (4% yield). Thus, acids, and the surface
51
despite not being active for hydrogenation oxygen behaves like Lewis bases. Thus, the
of the phenol ring, this catalyst was found activation of
to be active for deoxygenation. phenol on the oxide can take place through
Calculating the kinetic constant for this heterolytic dissociation of the OH bond on
reaction (corresponding to k2) further the oxide where the hydrogen from the phenol
is adsorbed on an oxygen site in the
showed that the deoxygenation capability oxide surface layer and a metal vacancy site
of this catalyst was on the same order of stabilizes the phenoxide ion. Generally,
magnitude as those of the other nickel alcohols have been found to almost always
catalysts, as shown in Table 5.
adsorb through heterolytic dissociation on
Hence, all catalysts supported on oxides
metal
are active for both hydrogenation and
51 54
oxides. Supporting this, Popov et al. obtained by hydrogenation of the aromatic ring, as
observed phenoxide formation on Al2O3 when described in
studying adsorption of phenol via IR section 3. Hydrogenation of an aromatic
55
spectroscopy. Liu et al. concluded that when hydrocarbon is, however, not trivial as these are
phenol is adsorbed at a Lewis acid site, one of considered to bind weakly to metal catalysts
58,59
the carbons in the aromatic ring becomes compared to conventional alkenes, but
highly nucleophilic and thereby receptive to adsorption on Lewis acid sites on the support
electrophiles, such as H+. Thus, the phenoxide appears to aid in this task.
will interact with a nickel crystallite where
hydrogen has been adsorbed, thereby 5. CONCLUSIONS
facilitating the saturation of the double bounds Screening experiments have been performed to
and producing cyclohexanone. These steps are identify ecient catalysts for HDO of phenol at
all illustrated in Figure 8. The cooperative intermediate temperature and pressure (275 C
eect of the support and active metal has also and 100 bar) in a batch reactor. Four groups of
been found for the phenol hydrogenation on catalysts were investigated: oxide catalysts,
palladium catalysts where the activity for methanol synthesis catalysts, reduced noble metal
hydrogenation can be signicantly increased catalysts, and reduced non-noble metal catalysts,
by the introduction of a Lewis acid because of totaling 23 catalytic systems.
the Across the series of investigated catalysts
5557
activation of the phenol on the Lewis sites. distinct dierences were observed in the catalytic
The formed cyclohexanol can directly react activity. Listing the apparent order of activity for
with the exposed metallic nickel surfaces as all the tested catalysts, the following is found:
shown in Figure 8 through adsorption at the
OH group. The actual deoxygenation step Ni/ZrO2 > NiV2O5/ZrO2 > NiV2O5/SiO2 >
probably takes place through a dehydration Ru/C
reaction, forming cyclohexene, as proposed in
the reaction scheme of Figure 3. The > Ni/Al2O3 > Ni/SiO2 Pd/C > Ni/MgAl2O4
cyclohexene can subsequently be
hydrogenated into cyclohexane as the nal > Ni/CeO2ZrO2 Ni/CeO2 > Pt/C
product. This reaction mechanism is more
complex than that discussed in the
Introduction and Figure 1b.
In summary, among all tested catalysts, only
those catalysts that showed both good
hydrogenation and deoxygenation capabilities
performed well, as a key step in the HDO of
phenol
under the tested conditions is the weakening
of the CO bond
cyclohexanone, which is rapidly Noble metal catalysts, in contrast to nickel,
hydrogenated to cyclohexanol. This can performed well on carbon supports, showing
that these metals on their own have the
then in a second step be dehydrated to
required anity for phenol activation probably
cyclohexene, by direct interaction with the aromatic ring.
which is rapidly hydrogenated to Ruthenium was found to be the most active
cyclohexane. The hydro- genation of the catalyst of the carbon-supported transition
aromatic ring causes a weakening of the metals, but overall, Ni/ZrO2 was the best
CO bond, making the deoxygenation
more favorable from cyclo- performing catalyst, with balanced rates of
hexanol than from phenol. both hydrogenation and deoxygenation. Thus,
A kinetic model revealed that the best nickel appears to be a promising, less
hydrogenation catalyst not necessarily expensive catalyst for HDO.
was the best deoxygenation catalyst. Pt/C,
Ni/ CeO2, and Ni/CeO2-ZrO2 were ASSOCIATED CONTENT
*S Supporting Information
found to be the best
performing hydrogenation catalysts, with their Derivation of the kinetic model, explanations
eectiveness of the assumptions made, and a nal
decreasing in that order, but were all validation of the model. This material is
available free of charge via the Internet at
found to have low activity for http:// pubs.acs.org.


deoxygenation.
Nickel was the best performing reduced AUTHOR INFORMATION
non-noble metal catalyst and was Corresponding Author
therefore tested on dierent supports. *E-mail: aj@kt.dtu.dk.
Oxide supports showed good activity, Notes
while a non-oxide support (as carbon) The authors declare no competing nancial interest.
showed no activity for phenol HDO.
However, Ni/C was active for HDO of
cyclohexanol, showing that the activation
ACKNOWLEDGMENTS
This work is part of the Combustion and Harmful
of the phenol prior to the hydrogenation Emission
Control (CHEC) research center at The
takes place on an oxygen vacancy site on
Department of Chemical and Biochemical
the oxidic support, but the subsequent
Engineering at the Technical
deoxygenation likely takes place directly nanced by
on the nickel crystallites.
Cu/ZnO/Al2O3 NiOMoO3/Al2O3 University of Denmark (DTU). This work is
DTU and The Catalysis for Sustainable Energy
NiCu/SiO2
initiative (CASE), funded by the Danish
> Cu/SiO2 > Co/SiO2 > Ni/C > Fe/SiO2, Ministry of Science, Technology and
MnO/C Innovation. Flash calculations were conducted
, WO3/C, V2O5/C, CoOMoO3/Al2O3 with SPECS version 5.62, which was kindly
supplied by the Center for Energy Resources
The tested oxide and methanol synthesis Engineering (CERE) at The Department of
catalysts had a low activity for the HDO of
Chemical and Biochemical Engineering of
phenol under the given conditions, which was
linked to their inability to hydrogenate the DTU. We nally thank Bodil Fliis Holten
phenol ring. Reduced metal catalysts of both (Center for Catalysis and Sustainable
noble and non-noble metals were signicantly Chemistry, Department of Chemistry, DTU) for
more active. On these catalysts, HDO of
phenol proceeds through an initial assistance with the measurements of the
hydrogenation to acidity of the catalysts.


(3) Holmgren, J.; Marinageli, R.; Nair, P.; Elliott, D. C.;
Bain, R.
REFERENCES Hydrocarbon Process. 2008, 95103.
(1) Mortensen, P. M.; Grunwaldt, J.-D.; Jensen, P. A.; (4) Raelt, K.; Henrich, E.; Koegel, A.; Stahl, R.;
Knudsen, K. G.; Jensen, A. D. Appl. Catal., A 2011, Steinhardt, J.; Weirich, F. Appl. Biochem. Biotechnol. 2006,
407, 119. 129, 153164.
(5) Zhang, W.; Zhan, Y.; Zhao, L.; Wei, W. Energy Fuels
(2)Bridgwater, A. V.; Czernik, S.; Diebold, J.; Meier, 2010, 24,
D.; Oasmaa, A.; Peakocke, C.; Piskorz, J.; Radlein, D.
Fast Pyrolysis of Biomass: A Handbook; CPL Press: 20522059.
Newbury, U.K., 1999. (6) Garca, I.; Lopes, J. M.; Cerqueira, H. S.; Ribeiro, M.
F. Ind. Eng. Chem. Res. 2013, 52, 275287.
(7) Furimsky, E.; Massoth, F. E. Catal. Today 1999, 52,
381495. (8) Furimsky, E. Appl. Catal., A 2000, (33) Prasomsri, T.; Nimmanwudipong, T.; Roman-Leshkov, Y.
199, 144190.
(9) Moraes, M. S. A.; Migliorini, M. V.; Damasceno, Energy Environ. Sci. 2013, 6, 17321738.
F. C.; Georges, F.; Almeida, S.; Zini, C. A.; Jacques, R. (34) Adriyanti, A. R.; Khromova, S. A.; Venderbosch, R.
A.; Caramao, E. B. J. Anal. Appl. Pyrolysis 2012, 98, H.; Yakovlev, V. A.; Heeres, H. J. Appl. Catal., B 2012,
5164. 117118, 105 117.
(10) Moberg, D. R.; Thibodeau, T. J.; Amar, F. (35) Elliott, D. C.; Hart, T. R. Energy Fuels 2009, 23, 631637.
G.; Frederick, B. G. J. Phys. Chem. C 2010, 114,
1378213795. (36) Davda, R. R.; Shabaker, J. W.; Huber, G. W.;
(11) Yakovlev, V. A.; Khromova, S. A.; Cortright, R. D.; Dumesic, J. A. Appl. Catal., B 2005,
56, 171186.
Sherstyuk, O. V.; Dundich, V. (37) Wildschut, J.; Mahfud, F. H.; Venderbosch, R. H.; Heeres, H.
J.
O.; Ermakov, D. Y.; Novopashina, V. M.; Lebedev, M.
Y.; Bulavchenko, O.; Parmon, V. N. Catal. Today 2009, Ind. Eng. Chem. Res. 2009, 48, 1032410334.
144, 362366.
(12) Vargas, A.; Burgi, T.; Baiker, A. J. Catal. 2004, (38) Lee, C. R.; Yoon, J. S.; Suh, Y.-W.; Choi, J.-W.; Ha, J.-
M.; Suh, D. J.; Park, Y.-K. Catal. Commun. 2012, 17,
222, 439449. 5458.
(13) Chorkendor, I.; Niemantsverdriet, J. W. (39) Adriyanti, A. R.; Gutierrez, A.; Honkela, M. L.;
Concepts of Modern Krause, A. O. I.; Heeres, H. J. Appl. Catal., A 2011,
407, 5666.
Catalysis and Kinetics; John Wiley & Sons, Inc.: New (40) Nrskov, J. K.; Rossmeisl, J.; Logadottir, A.;
York, 2007. Lindqvist, L.; Kitchin, J. R.; Bligaard, T.; Jonsson, H.
J. Phys. Chem. B 2004, 108, 1788617892.
(14)Chinchen, G. C.; Denny, P. J.; Parker, D. G.; (41) Foster, A. J.; Do, P. T. M.; Lobo, R. F. Top. Catal. 2012, 55,
Spencer, M. S.; Waugh, K. C.; Whan, D. A. Appl. 118128.
Catal. 1987, 30, 333338.
(15) Rasmussen, P. B.; Holmblad, P. M.; (42) Zielinski, J. J. Chem. Soc., Faraday Trans. 1997, 93,
Askgaard, T.; Ovesen, C. V.; Stoltze, P.; Nrskov, J. 35773580.
K.; Chorkendor, I. Catal. Lett. 1994, 26, 373 381. (43) Dry, M. E. J. Chem. Technol. Biotechnol. 2001,
(16) Askgaard, T.; Nrskov, J. K.; Ovesen, C. V.; 77, 4350. (44) Dry, M. E. Catal. Today 2002,
Stoltze, P. J. 71, 227241.
(45) Dry, M. E. FT Catalysts. In FischerTropsch Technology;
Catalysis 1995, 156, 229242. Steynberg, A., Dry, M., Eds.; Elsevier: Amsterdam,
(17)Vargas, A.; Reimann, S.; Diezi, S.; Mallat, T.; 2004; Chapter 7, pp 533593.
Baiker, A. J. Mol. Catal. 2008, 282, 18. (46) Rostrup-Nielsen, J. R. Steam Reforming. Handbook of
(18) Stakheev, A. Y.; Kustov, L. M. Appl. Catal., A Heterogeneous Catalysis; John Wiley & Sons, Inc.:
1999, 188, 335. New York, 2008; Chapter 13.11, pp 28822905.
(19) Baltes, C.; Vukojevic, S.; Schuth, F. J. Catal. (47) Filley, J.; Roth, C. J. Mol. Catal. 1999, 139, 245252.
2008, 258, 334 (48) Burwell, R. L.; Littlewood, A. B.; Cardew,
M.; Pass, G.; Stodhart, C. T. H. J. Am. Chem. Soc.
344. 1960, 82, 62726280.
(49) Burwell, R. L.; Taylor, K. C.; Haller, G. L. J. Phys.
(20) de Miguel Mercader, F.; Koehorst, P. J. J.; Chem. 1967,
Heeres, H. J.; Kersten, S. R. A.; Hogendoorn, J. A.
AIChE J. 2011, 57, 31603170. 71, 45804581.
(21) Schofield, K. Prog. Energy Combust. Sci. 2008, 34, (50) Rodriguez, J. A.; Hanson, J. C.; Frenkel,
330350. A. I.; Kim, J. Y.; Perez, M. J. Am. Chem. Soc. 2002,
124, 346354.
(22)Llewellyn, P. L.; Bloch, E.; Bourelly, S. Surface (51) Kung, H. H. Surface Coordinative Unsaturation.
Area/Porosity, Adsorption, Diusion. In Transition
Characterization of Solid Material and Heterogenous
Catalysts; Che, M., Vedrine, J. C., Eds.; Wiley- Metal Oxides: Surface Chemistry and Catalysis; Elsevier:
Amsterdam, 1989; Chapter 4, pp 5371.
VCH: Weinheim, Germany, 2012; Chapter 19, pp (52) Rethwisch, D. G.; Dumesic, J. A. Langmuir 1986, 2,
853880. 7379.
(23) Behrens, M.; Schlogl, R. X-ray Diraction and
Small Angle X-ray (53) Idriss, H.; Barteau, M. A. Adv. Catal. 2000, 45,
Scattering. In Characterization of Solid Material and 261331.
Heterogenous Catalysts; Che, M., Vedrine, J. C., (54) Popov, A.; Kondratieva, E.; Gilson, J.-P.;
Eds.; Wiley-VCH: Weinheim, Germany, 2012; Mariey, L.; Travert, A.; Mauge, F. Catal. Today 2011,
Chapter 15, pp 611654. 172, 132135.
(55) Liu, H.; Jiang, T.; Han, B.; Liand, S.; Zhou, Y. Science
(24) Shuikin, N.; Erivanskaya, L. Russ. Chem. Rev. 2009, 326,
1960, 29, 309320. 12501252.
(25) Zhao, C.; Kuo, Y.; Lemonidou, A. A.; Li, X.; (56) Velu, S.; Kapoor, M. P.; Inagaki, S.; Suzuki, K. Appl.
Lercher, J. A. Catal., A
Angew. Chem., Int. Ed. 2009, 48, 39873990. 2003, 245, 317331.
(26) Zhao, C.; He, J.; Lemonidou, A. A.; Li, X.;
Lercher, J. A. J. Catal. (57) Matos, J.; Corma, A. Appl. Catal., A 2011, 404, 103112.
2011, 280, 816. (58) Boitiaux, J. P.; Cosyns, J.; Robert, E. Appl. Catal. 1987, 32,
(27) Ohta, H.; Kobayashi, H.; Hara, K.; Fukuoka, A. 145
Chem. Commun. 168.
2011, 47, 1220912211. (59) Ponec, P.; Bond, G. C. Catalytic
(28) Zhao, C.; Kasakov, S.; He, J.; Lercher, J. hydrogenation and dehydrogenation. Catalysis by
A. J. Catal. 2012, 296, 1223. Metals and Alloys; Elsevier: Amsterdam, 1995; Vol. 95,
(29) Ryymin, E.-M.; Honkela, M. L.; Viljava, T.-R.; Chapter 11, pp 477539.
Krause, A. O. I.
Appl. Catal., A 2010, 389, 114121.
(30) Venderbosch, R. H.; Ardiyanti, A. R.;
Wildschut, J.; Oasmaa, A.; Heeres, H. J. J. Chem.
Technol. Biotechnol. 2010, 85, 674686.
(31) Benson, S. W. Thermochemical Kinetics: Methods for
estimation of
thermochemical data and rate parameters; John Wiley &
Sons, Inc.: New York, 1968.
(32) Fogler, H. S. Elements of Chemical Reaction
Engineering; Prentice
Hall: Upper Saddle River, NJ, 2006.

Você também pode gostar