Вы находитесь на странице: 1из 11

Journal of Catalysis 311 (2014) 1727

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Kinetic studies of hydrodeoxygenation of 2-methyltetrahydrofuran


on a Ni2P/SiO2 catalyst at medium pressure
Ayako Iino a, Ara Cho a, Atsushi Takagaki a, Ryuji Kikuchi a, S. Ted Oyama a,b,
a
b

Department of Chemical Systems Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
Department of Chemical Engineering, Virginia Tech, Blacksburg, VA 24061, United States

a r t i c l e

i n f o

Article history:
Received 27 March 2013
Revised 27 October 2013
Accepted 6 November 2013
Available online 8 December 2013
Keywords:
Nickel phosphide
Hydrodeoxygenation
2-Methyltetrahydrofuran
Reaction network
LangmuirHinshelwood kinetics

a b s t r a c t
Bio-oil obtained by the pyrolysis of woody biomass contains many oxygenated organic compounds which
degrade the product quality and make necessary upgrading for its use as a liquid fuel. Hydrodeoxygenation (HDO) is a catalytic hydrotreating process for the removal of the problematic oxygen functionalities
and is promising for bio-oil upgrading. In this work, 2-methyltetrahydrofuran (2-MTHF) was chosen as a
model oxygenated compound, and its HDO reaction mechanism was studied on a silica-supported nickel
phosphide catalyst (Ni2P/SiO2) at a medium pressure of 0.5 MPa. The temperature dependency of the
catalyst activity was determined and it was found that at 350 C Ni2P/SiO2 showed 100% conversion
and 85% selectivity to n-pentane, with higher oxygen removal activity and less CAC bond cracking activity than commercial noble metal Ru/C and Pd/Al2O3 catalysts based on the same amount of active sites. A
contact time study allowed the determination of a reaction sequence for 2-MTHF HDO on Ni2P/SiO2 and it
was found that CAO bond cleavage of the furanic ring to generate either 2-pentanone or 1-pentanal was
the rate-determining step. This was followed by hydrogen transfer steps to produce oxygen free
compounds, n-pentane or n-butane. A partial pressure analysis of 2-MTHF and H2 was consistent with
a rate equation derived using a LangmuirHinshelwood (LH) mechanism. This suggested that adsorption
of 2-MTHF and hydrogen occurred competitively and that these species reacted on the Ni2P/SiO2 surface.
Although high partial pressure of H2 was favorable for hydrogenation, too much H2 competed with
2-MTHF adsorption, which caused lower conversion.
2013 Elsevier Inc. All rights reserved.

1. Introduction
In recent times, biomass has been attracting much attention as
a sustainable resource with advantages over fossil fuels such as
renewability, lower expense, and a short CO2 cycle [13]. There
are a number of upgrading methods that are being considered to
transform biomass into usable form while increasing its energy
density [4]. Pyrolysis is one of these techniques, which consists
of the rapid heating of lignocellulosic biomass in an inert gas to
produce a liquid known as pyrolysis oil or bio-oil. Bio-oil is a promising alternative liquid fuel but contains considerable amounts of
oxygenated compounds derived from the lignocellulose feed such
as phenols, furans, carboxylic acids, ethers, and aromatic alcohols.
The oxygenated species in bio-oil amount to about 40 wt% oxygen
[5,6], which is a serious problem because oxygen compounds degrade the product quality, giving low pH, low heating value, high
viscosity, low thermal and chemical stability, and poor miscibility
with hydrocarbon fuels [79]. A method of upgrading the pyrolysis
Corresponding author.
E-mail address: oyama@vt.edu (S. Ted Oyama).
0021-9517/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcat.2013.11.002

oil is hydrodeoxygenation (HDO), the catalytic removal of oxygen


using hydrogen.
A number of catalysts have been developed and investigated for
HDO [10]. Among them are sulde catalysts such as CoMoS or NiMoS used in hydrodesulfurization (HDS) and hydrodenitrogenation
(HDN) processes to remove sulfur and nitrogen from conventional
petroleum feedstocks. Although pyrolysis oil does not have high
levels of these elements, which is an advantage, the removal of
oxygen is related to the removal of sulfur and nitrogen so similar
sulde catalysts were expected to be effective [1116]. It has been
found that these catalysts have reasonable activity for the HDO of
many kinds of oxygen functionalities such as carbonyl, carboxylic,
guaiacyl, guaiacolic [11,13], and phenolic groups [12]. However, a
fundamental problem with the use of sulde catalysts in HDO is
that to be active these catalysts must be in a sulded form, so
sulfur compounds must be added to the feed to produce a sulding
atmosphere [15]. For this reason, non-sulde catalysts are also
been investigated. Noble metal catalysts such as Pt, Pd, Rh, and
Ru are gathering attention as highly active HDO catalysts
[1720]. A study on zirconia-supported mono- and bimetallic
noble metal (Pt, Pd, Rh) catalysts has shown that these catalysts

18

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

are more active for guaiacol HDO and cause less carbon deposition
than a conventional sulded CoMo/Al2O3 catalyst [17]. The catalytic performance of platinum supported on a mesoporous zeolite
in benzofuran HDO has been examined, and it has been shown that
platinum supported on mesoporous ZSM-5 has better activity than
on conventional ZSM-5 or Al2O3 because of its combination of high
acidity and easily accessible channels [18]. The aqueous-phase
HDO of phenols over noble metal catalysts has been studied, and
it is reported that a balance between metallic and acidic functions
in combination results in good catalytic activity [19]. The aqueousphase HDO of carboxylic acids was studied over supported Ru catalysts by kinetic and diffuse reectance infrared measurements
and it was found that the hydrogenation of C@O bonds was favored
by acidic supports while the cleavage of CAC bonds was favored in
the order of Ru/C > Ru/ZrO2 > Ru/Al2O3 [20].
In recent decades, transition metal phosphide catalysts have attracted great attention for HDS and HDN [2123] and also HDO
[24,25]. The subject has been recently reviewed [23]. In particular,
in previous studies, it was shown that nickel phosphide catalysts
had higher activity for HDS and HDN than other phosphides [22].
Recently, Ni2P has also been shown to have higher activity for
the HDO of anisole [26], guaiacol [27], and 2-methyltetrahydrofuran [2,28] than other transition metal phosphide catalysts like
Co2P, Fe2P, WP, and MoP. Furthermore, it was reported that Ni2P/
SiO2 has bifunctional acid and metal functions which resulted in
high performance in hydrotreating process [29,30]. The activity
of the catalysts is affected by the support material [31] and the
method of synthesis [32,33].
In this work, a Ni2P/SiO2 catalyst was prepared and characterized for the HDO of 2-methyltetrahydrofuran (2-MTHF) a model
species derived from furan, which is a substantial component of
pyrolysis oil [5]. First, it is shown that the activity of the Ni2P/
SiO2 compares favorably with those of commercial Pd/Al2O3, and
Ru/C, which are known to have good activity for HDO [20]. Then,
the reaction network in the HDO of 2-MTHF was determined by
contact time measurements. Furthermore, the effect of 2-MTHF
or H2 partial pressure on the catalyst activity was tested and discussed. Pyrolysis oil is a complex mixture with a large number of
oxygenated compounds, so the study of the hydrodeoxygenation
of pyrolysis oil is very challenging. For this reason, the study of
reaction mechanisms of model oxygenated compounds is important to provide basic information about removal of oxygen from
pyrolysis oil. It has been reported that the HDO of furans is more
difcult than that of other oxygen-containing compounds like
carboxylic acids, ketones, and phenols [34]. The model compound
2-methyltetrahydrofuran (2-MTHF) was chosen for study because
the presence of the methyl group (Fig. 1) permits studies of reaction pathway as ring-opening can occur in two directions, an open
and a hindered side.
Previous work on the HDO of 2-MTHF in Ni2P/SiO2 [2] differs
substantially from that reported here. The previous work was at
0.1 MPa, which resulted in a reaction sequence in which olens
were primary products. The present work is at 0.5 MPa, which
causes a different reaction sequence in which 2-pentanone and
n-pentanal are primary products. The previous study employed
XPS to conrm the presence of phosphide phases prepared in
two different ways, which the present study uses XPS to show that

Fig. 1. Structure
possibilities.

of

2-methyltetrahydrofuran

(2-MTHF)

and

ring-opening

carbon is not deposited after reaction. The rst study did not involve reaction network analysis, while the present work presents
a modeling study which indicates that ring-opening is the ratedetermining step. The earlier paper did not conduct reaction kinetic studies, while the present work presents a partial pressure
analysis with derivation of a rate expression. The rate expression
is based on a sequence of steps with a surface that is sparsely occupied by reaction products, which is consistent with the nding that
conversion varies linearly with contact time. Thus, the objectives of
this study are to report on a novel, stable catalyst with high selectivity to desired HDO products at moderate temperature and pressure, to explore the reaction sequence for HDO of 2-MTHF at these
conditions, and to provide insight on the state of the surface during
reaction.

2. Experimental methods
A fumed silica EH-5 provided by Cabot Corp. was employed as a
support material. Chemicals used were Ni(NO3)26H2O (SigmaAldrich, 99.999%), (NH4)2HPO4 (SigmaAldrich,>99.0%), HNO3 (TCI,
67%). Samples of 5 wt% Pd/Al2O3 and 5 wt% Ru/C were provided
by BASF Catalysts. The chemicals used as reactants were 2-methyltetrahydrofuran (SigmaAldrich,>99.0%) and n-heptane (TCI,>
99.0%). The chemicals used for calibration of the gas chromatograph were n-butane, 2-methylfuran (TCI,>98.0%), 2-pentanone
(TCI,>97.0%), furan (TCI,>99.0%), n-pentanal (TCI,>95.0%), n-pentane, 1-pentene, trans-2-pentene, 1-pentanol (SigmaAldrich,
99%), cis-2-pentene, and 2-pentanol (SigmaAldrich, 98%). Bulk
Ni2P (SigmaAldrich, 98%) was used for measurement of reference
XRD patterns.
2.1. Preparation of catalysts
The Ni2P/SiO2 catalyst was prepared by a temperature-programmed reduction method from supported phosphate precursors.
First, the dried silica support was impregnated to the incipient
wetness point with an aqueous solution containing Ni(NO3)26H2O
and (NH4)2HPO4, the amount of Ni was 1.156 mmol/g-support
(7.9 wt%), and the initial ratio of Ni/P was 1/2. Second, a phosphate
precursor for the phosphide was obtained after drying at 112 C for
12 h calcining at 500 C for 6 h in ambient air, and then pelletizing
and sieving to a size of 6501180 lm. Third, the precursor was reduced from room temperature to 590 C at a rate of 3 C/min and
maintained at 590 C for 2 h. The H2 (>99.9999%) ow was set at
1000 mL/min per gram of precursor. The prepared catalyst was
cooled to room temperature in a He ow then passivated in a
0.5 vol% O2/He ow (100 mL/min) for 4 h.
2.2. Characterization of catalysts
The reducibility of precursors was characterized by temperature-programmed reduction (TPR) using a quartz reactor, in which
around 0.1 g of catalyst was loaded. After drying at 150 C for 2 h,
reduction was conducted at a heating rate of 3 C/min in H2
(1000 mL/min per g precursor). The H2O production was monitored by a quadrupole mass spectrometer.
The quantity of active sites was estimated by pulse injection of
CO using 0.10.3 g quantities of catalysts loaded in a U-tube quartz
reactor and reduced in H2 (100 mL/min) at 450 C for 1 h. Then,
with owing He (100 mL/min), pulses of CO (25 lL) were injected
sequentially at room temperature. Outlet gases were detected by a
thermal conductivity detector (TCD).
The crystal structure of the catalysts was determined by X-ray
diffraction (XRD) analysis using a Rigaku XRD diffractometer

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

operated with Cu Ka monochromatized radiation generated at


40 kV and 100 mA.
BET surface areas and pore volumes were obtained from the linear portion of BET plots at liquid nitrogen temperature using a
Micromeritics ASAP 2010. Samples were dried at 120 C and evacuated for 1 h, before measurement.
The near-surface properties of the catalysts were measured by
X-ray photoelectron spectroscopy (XPS) measurements carried
out using a PHI 5000 Versaprobe with a Mg Ka X-ray source
(1486.7 eV) and a spherical section analyzer. The XPS spectra were
collected using a pass energy of 23.5 eV. For XPS measurement, the
passivated catalyst was ground to powders with a mortar and pestle, and pelletized by pressurizing to 20 MPa for 10 min.
2.3. Catalyst reactivity measurements
Reaction was carried out in a continuous-ow quartz reactor
(inner diameter of 0.8 cm and total length of 28 cm), with the catalyst loaded in a section about 2 cm long in the middle of the reactor with quartz sand of the same particle size to disperse the
catalyst and ensure uniform ow. The passivated Ni2P/SiO2 catalyst
was pretreated in a H2 ow (150 mL min1) at 450 C for 2 h (Ni2P/
SiO2), at 150 C for 1 h (Pd/Al2O3) and at 200 C for 1 h (Ru/C) at
atmospheric pressure. A mixture of 95 vol% of 2-MTHF and
5 vol% of n-heptane as an internal standard was vaporized and
mixed with a H2 gas stream to give a reactant stream of 3 or
5 mol% 2-MTHF in H2. For most measurements, the total pressure
was xed at 0.5 MPa using a back-pressure regulator. A Shimadzu
gas chromatograph GC-14A (GC) equipped with a ame ionization
detector (FID) and a thermal conductivity detector (TCD) and a
commercial HP-1 capillary column (100 m  0.250 mm  0.50 lm)
was used to analyze the products.
The conversion of 2-MTHF (XMTHF), total HDO percent (XHDO),
and selectivity (Si) were dened based on the moles of each chemical as follows:



nMTHF
 100%
X MTHF 1 
nMTHF;0
P
X HDO

hydrocarbons Carbon number

Contact time s

Quantity of activ e sites lmol=g  Catalyst weight g


Reactants flowrate lmol=s
6

The amount of active sites was measured by CO chemisorption.


The H2 ow rate was adjusted to obtain a molar ratio of 2MTHF:H2 to 5:95.
To determine the effect of reactants partial pressure, reaction
experiments changing H2 partial pressure or 2-MTHF partial pressure were conducted using N2 as the balance. A quantity of 0.05 g
of Ni2P/SiO2 was used with 1 g of quartz sand, and the total pressure and the temperature were xed at 500 kPa (0.5 MPa) and
300 C. To check the H2 partial pressure effect, the 2-MTHF ow
rate and partial pressure were xed 3 lmol/s and 9 kPa and the
H2 partial pressure was changed in the order; 418, 332, 142, 75,
95, 237 kPa by changing H2 ow rate. To check the 2-MTHF partial
pressure effect, H2 ow rate and partial pressure were xed at
166 mL/min and 332 kPa, and 2-MTHF partial pressure was changed in the order; 6, 15, 39, 24, 9 kPa. N2 was used for balance to
keep the total pressure at 500 kPa.
3. Results and discussion

 ni

5  nMTHF;0

!
 100%


ni
 100%
nMTHF;0  nMTHF

where nMTHF,0 and nMTHF express the moles of 2-MTHF in the feed
and product, respectively, and ni is the moles of product i. Total
HDO percentages show how much 2-MTHF is converted into hydrocarbons without oxygen like n-pentane, n-butane, and pentenes.The
product distribution of species i (PDi) was dened based on the carbon number of each species as follows:



Carbon number  ni
 100%
PDi P
Carbon number  ni

The carbon balance (CB) was calculated from:

CB

350 C and was varied from high to low and up again to verify that
the reactivity data obtained was for a stable catalyst. The temperature was altered in the order: 350, 300, 200, 250, 325, and 275 C
for Ni2P/SiO2 and Pd/Al2O3, 350, 300, 200, 250, 150, and 275 C for
Ru/C, with each condition maintained for 34 h. For post-reaction
testing, the catalyst was ushed with N2 for 1 h and then removed
from the reactor for characterization by XRD, CO chemisorption, N2
physisorption, and XPS. A similar procedure was used for stability
testing.
For contact time analysis, small amounts of the Ni2P/SiO2 catalyst (0.030, 0.050, 0.075 g) and 1 g of quartz sand were mixed and
loaded in the reactor. After pretreatment, the pressure and the
temperature were xed at 0.5 MPa and 300 C. The products were
analyzed by GC at various reactant ow rates. The contact time was
dened as follows:


Si

19

P

Carbon number  ni
 100%
5  nMTHF;0

This value was used to determine the carbon loss during reactions.
For activity testing, amounts of catalysts corresponding to
16.5 lmol of CO uptake sites were used, so quantities of 0.30 g of
Ni2P/SiO2, 0.21 g of Pd/Al2O3, and 0.12 g of Ru/C were loaded with
1 g of quartz sand in the reactor. After pretreatment, the pressure
was xed at 0.5 MPa and the liquid reactants ow rate was set at
3.0 lmol/s and the H2 ow rate was adjusted to obtain a molar ratio of 2-MTHF:H2 of 3:97. The temperature was slowly raised to

3.1. Catalysts characterization


Table 1 shows the surface properties of the various catalysts.
The number of CO uptake sites of Ni2P/SiO2 is 55 lmol/g, of 5%
Pd/Al2O3 80 lmol/g, and of 5% Ru/C 140 lmol/g. Particle sizes are
calculated from the CO uptakes, assuming spherical particles as detailed in the Supplementary information section, Part 1. For Ni2P/
SiO2, the number of CO uptake sites of the spent catalyst which
was used for the catalyst stability test was the same as for the fresh
catalyst. This result indicates that the structure of Ni2P/SiO2 does
not change appreciably during the reaction. The BET surface area
of the SiO2 support was 273 m2/g and that of the fresh Ni2P/SiO2
and the spent Ni2P/SiO2 were 214 and 220 m2/g, respectively.
The reason the surface area of Ni2P/SiO2 is smaller than that of
the SiO2 support is that the intraparticle pores of SiO2 are partially
lled with the impregnated catalyst phase. The BET surface area
also did not change signicantly before or after the reaction, conrming that the surface structure of the catalyst did not change
in the reaction.
Fig. 2 shows TPR proles for mass 18 (H2O) and mass 34 (PH3) for
the synthesis of Ni2P/SiO2 from the nickel phosphate precursor. The
H2O signal shows a peak centered at 590 C for reduction of the metal, while the PH3 signal shows a corresponding peak at 598 C for
reduction of the phosphate, but with a tail to higher temperature.
The tail occurs because the precursor contains excess phosphate
and PH3 is generated from its reduction beyond the temperature
at which the phosphate is reduced to the corresponding metal.

20

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

Table 1
Surface properties of supports and catalysts.
CO uptakeb
(lmol/g)

BET surface area


(m2/g)

Particle size
(nm)

SiO2 (EH-5)
Fresh Ni2P/SiO2
Spent Ni2P/SiO2a
5% Pd/Al2O3
5% Ru/C

55
55
80
140

273
214
220
82
Proof stage

22
22
6.5
3.6

Spent Ni2P/SiO2

300 C for 11 h.
After in situ reduction (H2 100 mL/min for 1 h at 450 C).

Intensity / A.U.

a
b

Catalyst

Mass = 18 (H2O) Intensity / a.u.

590

Fresh Ni2P/SiO2

SiO2

Bulk Ni2P
10

200

400

600

Mass = 34 (PH3) Intensity / a.u.

600

40

50

60

70

80

Fig. 3. XRD patterns for Ni2P/SiO2, support silica and bulk Ni2P.

598

400

30

800

Temperature /C

200

20

1/5

800

Temperature /C
Fig. 2. TPR proles of precursor of Ni2P/SiO2.

From these results, the reduction temperature for the larger scale
catalyst syntheses was chosen to be 590 C.
Fig. 3 shows XRD patterns for the spent and fresh Ni2P/SiO2, the
support SiO2, and a reference bulk Ni2P sample. The intensity of the
pattern for the bulk Ni2P has been reduced by a factor of ve because it is a pure sample with a strong signal. The XRD pattern
for silica shows a broad peak at around 22 characteristic of amorphous silica. The pattern for fresh Ni2P/SiO2 shows peaks at around
41, 44, 47, and 55, which line up with the peaks of bulk Ni2P.
The peaks have low intensity and are wide, indicating that small
Ni2P crystallites are formed that are highly dispersed on the silica
support. The pattern for spent Ni2P/SiO2 shows similar peaks, indicating the crystal structure of Ni2P/SiO2 is not altered during the
reaction.
Fig. 4 shows the XPS spectra for Ni 2p and P 2p lines of the
freshly reduced and passivated catalyst and spent Ni2P/SiO2

catalyst. The spectra were taken with samples exposed to the


atmosphere, and the main reason for the analysis was to ascertain
that no excess carbon was present on the spent material. There is
little difference in the two spectra. For Ni 2p3/2, two main peaks appear at 856.8 eV and 852.4 eV with corresponding Ni 2p1/2 peaks at
869.7 eV and 873.5 eV. For P 2p, the peak appears at 134.4 eV. Previous studies have shown that the Ni 2p3/2 peak at 856.2 eV can be
assigned to Ni2+ and the 853.1 eV peak to Nid+ (0 < d < 2) in Ni2P,
while the P 2p peak at 133.0 eV can be assigned to P5+ and the
129.5 eV peak to Pd (0 < d < 1) in Ni2P [35]. The peaks representing
Ni2+, Nid+ and P5+ can be seen but the peak of Pd cannot be seen
clearly. This is likely because small Ni2P crystals on silica are oxidized due to exposure of the samples to air. For C 1s, the peaks appear at 284.4 eV for the both fresh and spent catalysts and the
element ratio calculated by the peak area and the photoionization
cross-sections of Si 2p and C 1s are C/Si = 0.11 for the fresh catalyst
and C/Si = 0.10 for the spent catalyst. The reason the C 1s peak appears on the fresh catalyst is that carbon from contaminants in air
is deposited on the catalyst surface. This suggests that the carbon
on the surface of the spent catalyst is also just contamination from
air and carbon deposition does not occur to a substantial degree
during the reaction.
3.2. Reactivity measurements
3.2.1. Catalysts stability test
Fig. 5 shows conversion of 2-MTHF and product distributions on
Ni2P/SiO2 at 0.5 MPa, 300 C during 11 h on stream. Methane, ethane, and propane were not produced, and other species like pentenes and pentanal were generated at levels under 0.1%. The carbon
balance was 100% 5%. The uncertainty in the carbon balance does
not affect the calculation of the selectivity signicantly as discussed in the Supplementary information, Part 2. Measurements
were conducted for 11 h during which time the conversion

21

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

(a) Ni 2p

856.4
873.3

Intensity / a.u.

869.7

853.0

Spent Ni2P/SiO2

856.8
852.4
Fresh Ni2P/SiO2

873.5
869.7

885

880

875

870

865

860

855

850

845

Binding energy / eV

(b) P 2p

134.4

Intensity / a.u.

134.4
129.5

Spent Ni2P/SiO2

129.5
Fresh Ni2P/SiO2

142

140

138

136

134

132

130

128

126

124

Binding energy / eV
Fig. 4. XPS spectra of the fresh and spent Ni2P/SiO2 catalyst (a) shows the spectra
for Ni 2p and (b) for P 2p.

decreased from 63% to 50% and then stabilized, with the product
distribution not changing signicantly. The main product was npentane (80%), followed by n-butane (14%) with small amounts
of CO (3%), 2-pentanone (2%), 1-pentanol (0.5%), and 2-pentanol
(under 0.2%). The amount of CO was always around 1/4 of that of
n-butane. This indicates that 2-MTHF, which has ve carbons, is
converted into n-butane and CO, because product distributions
are calculated based on carbon number of each product. This
means that the same moles of CO and n-butane were produced,
which indicates some C5 oxygenated organic compounds are converted into n-butane and CO through CAC bond cleavage.

(a)

n-Pentane

80

60

Conversion

100

Conversion Total HDO


Ru/C

20

n-Butane

0
30

100

200

300

400

500

CO

(b)

600

2-pentanone

1-Pentanol

2-Pentanol

100

200

300

400

500

600

Reaction time / min

Conversion, Total HDO / %

40

Conversion, Product
distribution / %

Conversion,
Product distribution / %

100

3.2.2. Catalytic activity test


The n-heptane internal standard was unreactive at all conditions for Ni2P/SiO2 and Pd/Al2O3 and for Ru/C below 250 C. Cracking of n-heptane was only observed at 275 C and above on Ru/C at
conversions above 80% but this did not affect the conclusions on
reactivity. The carbon balance on Ru/C at 150 and 200 C and on
Ni2P/SiO2 and Pd/Al2O3 at all temperatures are in the range of
100% 5%. Although pentenes and pentanal and 2-methylfuran
were also detected besides the products shown here, they were
generated at under 0.1% so the product distributions of these
chemicals are not shown. A blank activity test using pelletized
and sieved support silica instead of the catalysts was conducted
and at 350 C the conversion of 2-MTHF was 0.5% at most. So,
the HDO activity of the silica support is negligible.
Fig. 6 shows conversions and total HDO percents of 2-MTHF for
Ni2P/SiO2, Pd/Al2O3, and Ru/C, which as expected, rise with increasing temperature and reach almost full conversion at 350 C.
Comparisons to the noble metals are appropriate because considerable work has been reported on Pd [36,37] and Ru [38,39] for
hydrodeoxygenation reactions. We acknowledge that the supports
are different, but the support effects are secondary to the action of
the metals. Selectivity as a function of temperature for each catalyst is reported in the Supplementary information, Part 3.
Conversions, product distributions, and carbon balances on
each catalyst at 250 and 350 C are summarized in Table 2. At
250 C, the conversion follows the order Ru/C (40%) > Pd/Al2O3
(20%) > Ni2P/SiO2 (6.2%) and the total HDO percent follows the order Ru/C (25%) > Pd/Al2O3 (6.6%) > Ni2P/SiO2 (2.6%). At 250 C, the
main product on Ru/C is methane (32%), on Pd/Al2O3 is 2-pentanol
(26%), and on Ni2P/SiO2 is n-pentane (67%). At 350 C on Ru/C, the
product detected is mostly methane (99%). This indicates the
occurrence of multiple hydrogenolysis. Even the n-heptane added
to the reactants as an internal standard is also completely converted into methane. The generation of alkanes with carbon number under 5, such as n-butane, propane, ethane and methane,
means that cracking of CAC bonds occurs. This CAC bond cracking
is easier on noble metal catalysts, especially on Ru/C than on Ni2P/
SiO2.
Fig. 7 shows the product distributions of each catalyst at temperatures when the conversion is around 15% (13% at 200 C on
Ru/C, 20% at 250 C on Pd/Al2O3, and 14% at 275 C on Ni2P/SiO2).
The selectivity to hydrocarbons increases in the order Ru/C < Pd/
Al2O3 < Ni2P/SiO2. On Ru/C, the main product is 2-pentanol (65%),
on Pd/Al2O3, similar amounts of n-pentane, 2-pentanone, and
pentanols are generated, while on Ni2P/SiO2, the main product is

80

Ni2P/SiO2
Pd/Al2O3

60

40

20

0
150

Fig. 5. (a) Conversion () of 2-MTHF and product distributions to the main
products, n-pentane, n-butane, CO, 2-pentanone, 1-pentanol and 2-pentanol in
elapsed time at 300 C, 0.5 MPa with 3% 2-MTHF in H2 and (b) expanded scale in the
range of 03.5%.

200

250

300

350

Temperature /C
Fig. 6. Conversion and total HDO percent on Ni2P/SiO2, Pd/Al2O3 and Ru/C at
0.5 MPa with 3% 2-MTHF in H2.

22

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

Table 2
Conversion, product distribution and carbon balance at 250 and 350 C at 0.5 MPa with 3% 2-MTHF in H2.

T (C)

Sample

Conv. (%)

Product distribution (%)

CB (%)

CO

Methane

Ethane

Propane

Butane

Pentane

2-Pentanone

2-Pentanol

1-Pentanol

250

Ni2P/SiO2
Pd/Al2O3
Ru/C

6.21
19.8
39.6

0
0
0

0
0.22
32

0
0
5.8

0
0.06
8.8

10
5.0
9.0

67
25
4.7

6.5
21
22

5.8
26
18

11
23
0

97
102
*

350

Ni2P/SiO2
Pd/Al2O3
Ru/C

99.8
98.7
100

2.8
0
0

0
3.0
99.4

0
0.06
0.52

0
0.64
0.05

12
24
0.01

85
71
0

0.072
1.3
0

0
0.08
0

0
0.18
0

103
97
*

No data, because of high cracking activity.

n-pentane (75%) On Ni2P/SiO2, oxygen removal occurs with little


generation of low molecular C1AC3 products by CAC cracking.
High selectivity to n-pentane is desirable, because the retention
of carbon in the product maximizes the heat content and quality
of the produced bio-oil [40]. The comparison here is at different
temperatures, because of the desirability of making comparisons
at the same conversion. A comparison at the same temperature
and conversion would be ideal, but would require changes in space
velocity, which would introduce distortions of its own in terms of
unequal productivities.
In our previous study at 0.1 atm [2], Pd/Al2O3 was found to have
a considerably lower activity than Ni2P/SiO2. In this study at
0.5 atm, the activity is comparable, and this may have to do with
a higher cleanliness of the Pd surface with a more elevated H2
pressure.

3.2.3. Contact time analysis


A plot of conversion versus contact time shows a linear relation
up to about 40% conversion (Fig. 8). An extra point (circle) was
added from analysis of data at different partial pressure (Section 3.2.6), but the same 2-MTHF/H2 partial pressure ratio (5/95),
which shows that there is no tendency for the line to curve downwards. Details of the calculation are in the Supplementary information, Last part.
The linearity indicates that within this range of conversion each
2-MTHF reactant molecule is reacting independently, meaning that
each reactant molecule sees essentially a fresh surface. By fresh
surface is meant a surface that is largely occupied by hydrogen
atoms with a few landing sites, and which looks equivalent to each
incoming 2-MTHF molecule. The implication is that there is no major coverage with other organic species, as if there were coverage,
the rate of adsorption would be retarded by those species as given
by the well-known expression, rads = k(reactant)(1  h). Thus, this
region of conversion represents differential conditions in kinetics,
and the resulting selectivity is devoid of complications due to
multiple reaction of products following readsorption. Selectivity
to the main products, n-pentane, n-butane, 2-pentanone, 1-pentanol, 2-pentanol, and 2-pentene versus contact time shows characteristic curves for each product (Fig. 9). Selectivity to other
products like 1-pentene and n-pentanal are under 0.1%, and results
for them are not shown. Carbon mass balances for these data are in
the range of 100% 5%. The selectivity to n-pentane increases
monotonically as contact time becomes longer. For n-butane, a
similar curve is seen though it becomes constant at around 20%.
On the contrary, the selectivity to 2-pentanone decreases as contact time becomes longer (Fig. 9 a). The selectivity to 1-pentanol
and 2-pentanol increases at shorter contact times and decreases
at longer contact times (Fig. 9b). From these results, the order of
formation of products can be deduced, and they suggest that
2-pentanone is a primary product, pentanols are secondary products and n-pentane and n-butane are nal products. There is also

a small amount of 2-pentene formed, whose selectivity increases


at low contact time (Fig. 9b). Although, this formally indicates it
is a primary product, it is very likely that because it is formed at
less than 1% levels the downturn of its selectivity is simply not
seen at these conditions.

3.2.4. Proposed reaction network


From the contact time analysis, a reaction network for the HDO
of 2-MTHF on Ni2P/SiO2 at 300 C, 0.5 MPa can be proposed
(Fig. 10). There are two main reaction routes. The rst route involves ring-opening of the tetrahydrofuran ring on the less-hindered side, where there are no substituents, and the second
route involves ring-opening on the more-hindered side, where
the methyl group is located. In the rst route, 2-pentanone is
formed initially, which is then converted into 2-pentanol. In turn,
2-pentanol is transformed into 2-pentene by dehydration, and
the 2-pentene becomes n-pentane by hydrogenation. The 2-pentanol can also form 1-pentene, but this is less favored because it is a
less-substituted alkene. The direct conversion of alcohols into
alkanes has been reported in some studies on Pd catalysts [41],
but seems to go by a more circuitous route on Ni2P. In the second
route, 2-MTHF is initially converted into pentanal, but the aldehyde is not seen as a product because it is rapidly converted into
other species due to its high reactivity. Pentanal reacts by two
pathways. In one pathway, pentanal forms 1-pentanol, a putative
intermediate formed in low concentrations, and then n-pentane
by dehydration and hydrogenation. In the other pathway, pentanal
reacts to form n-butane, probably by decarbonylation to form a CO
coproduct [20,42]. The formation of CO was conrmed in the catalyst activity tests. Methane, however, was not seen in the products
of Ni2P/SiO2, indicating that hydrogenation of the CO is slow at the
conditions of this study. In the contact time analysis experiments,
CO was not detected among the products because of the detection
limits of the TCD detector in the GC when the conversion of
2-MTHF is low. Another reason why CO production is low could
be carbon deposition from disproportionation. However, CO2 is
not observed and although carbon is detected on the surface of
the spent Ni2P/SiO2, it is not above the baseline level of the fresh
Ni2P/SiO2 by XPS analysis. The stability of the catalysts indicates
that carbon deposition likely does not occur at the reaction conditions employed. To conrm that n-butane is generated from pentanal and not from cracking of pentane, pentanal HDO was
conducted at the same conditions of 0.5 MPa and 300 C on Ni2P/
SiO2 and full conversion of pentanal and generation of n-butane
and CO were conrmed. It should be stressed that the network is
simplied and preliminary, as some of the steps may be reversible,
and some possible steps are omitted (e.g. formation of 1-pentene
from 2-pentanol).
The Ni2P catalyst has excess phosphorus, which resides on the
silica support as an acidic phosphate and likely aids in the adsorption of the 2-MTHF. The steps in the network involve hydrogen

23

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

80

n-Pentane

Oxygenates
2-C5OH

2-ONE

60

Hydrocarbons C1

Selectivity / %

Ru/C (200 C)
Conv. = 13.3 %

C3

C2 C4 C5

Oxygenates
2-ONE

1-C5OH

2-C5OH

Pd/Al2O3 (250 C)

40

n-Butane

20

2-Pentanone

Conv. = 19.8 %

Hydrocarbons

C4
C5
2-C5OH

0.0
4
Ni2P/SiO2 (275 oC)

1-C5OH
2-ONE

Conv. = 14.4 %

Hydrocarbons
C4

C5

25

50
75
Product distribution / %

100

Selectivity / %

Oxygenates

0.5

1.0

1.5

2.0

1.5

2.0

2-Pentanol

3
2

1-Pentanol

2-Pentene

0
0.0

0.5

1.0

Contact time / s
Fig. 7. Product distribution at various temperatures at 0.5 MPa and with 3% 2MTHF in H2 at a conversion of around 15%.

Fig. 9. Selectivity to the main products, n-pentane, n-butane, 2-petanone, 2pentanol, 1-pentanol, and 2-pentene at 300 C, 0.5 MPa with 5% 2-MTHF. (a)
Selectivity in the range of 080% and (b) Expanded scale in the range of 04%.

40
35

Conversion / %

30
25
20
15
10
5
0
0.0

0.5

1.0

1.5

2.0

2.5

Contact time / s
Fig. 8. Conversion of 2-MTHF on Ni2P/SiO2 as a function of contact time at 300 C,
0.5 MPa with 5% 2-MTHF in H2. The circle shows conversion calculated from rate
data at a different partial pressure.

transfer reactions that probably occur on the Ni2P. The results


obtained in this study at 0.5 MPa are to be contrasted with those
obtained earlier [2,29] on a similar catalyst at the same temperature. One study was carried out at 0.1 MPa [2], and at the lower
pressure, the primary products were alkenes, whereas in this study
the alkenes are not observed and the primary intermediates are
suggested to be either 2-pentanone or n-pentanal. The differences
will be discussed in the next section, and as will be seen can be
attributed to a change in mechanism due to differences in surface
coverage of adsorbed intermediates. Another study was carried out
at 0.5 MPa [29] as in this study, and a similar sequence was found.
However, in the previous study, the catalysts were a series of NiFeP
catalysts supported on KUSY and the objective was to probe the effect of ensemble effects. It was found that in the NiFeP alloys the
iron acted as an inert diluent, and the TOF based on surface Ni
atoms did not depend on composition, indicating that the reaction
was structure-insensitive, and that only one Ni atom was involved
in the rate-determining step.
3.2.5. Kinetic analysis
Fig. 11 shows the results of mathematical calculations for the
pseudo rst-order models of the proposed reaction network.

Reaction rate constants (k1k9) are dened for each reaction step
shown in Fig. 10. Because in this case, the amount of H2 was significantly larger than the amount of 2-MTHF, the amount of H2 was
assumed to be constant. Two intermediates in the sequence were
not observed (n-pentanal and 1-pentene) and to handle this in
the tting they were arbitrarily assigned very low concentrations.
The differential equations are presented in the Supplementary
information section, Part 4. To evaluate this calculation, the regression coefcient was determined and was found to have a value of
0.9, which shows a moderate good t.
From this analysis, it is deduced that the rate-controlling step of
the reaction network is the tetrahydrofuran ring-opening step, the
rst step of this reaction by CAO bond cleavage. Moreover, since
k5 > k1, unexpectedly the ring-opening on the more hindered side
is favored.
The direction of ring-opening can be understood from fundamentals of the reaction mechanism. For the top route of Fig. 10 that
forms 2-pentanone, the ring is likely opened by hydride attack on
the less hindered carbon in an SN2 type reaction (Fig. 12, top). For
the bottom route of Fig. 10 that forms n-pentanal, the ring is
opened by formation of a secondary carbonium ion in an SN1 type
reaction (Fig. 12, bottom). Thus, the preference for ring-opening on
the more hindered side can be ascribed to the limited basicity of
the hydride ion, so that the SN2 attack is slow, and also the stability
of the secondary carbonium ion, which permits the SN1 reaction to
proceed.
As mentioned in the previous section in a previous work [29]
with the same catalyst and reaction at a low pressure (0.1 MPa),
the primary intermediates were observed to be olens, while at
medium pressure (0.5 MPa), they were found to be 2-pentanone
or n-pentanal. This can be rationalized in the following manner
(Fig. 13). In both pressure regimes, the rst step is adsorption of
2-MTHF followed by ring-opening by hydride attack to form alkoxide intermediates (Fig. 12). At low-pressure conditions, the surface
is only partly occupied by adsorbed species and there are open
sites that can take up hydrogen atoms to produce the observed olens by dehydration (Fig 13a). The dehydration can be viewed as a
b-hydride elimination from a position two carbon atoms away
from the oxygen atom. In the medium-pressure regime the surface
is more fully occupied with adsorbed species, the organic intermediates are likely more erect on the surface, and reaction steps on

24

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

k1

k2

k3

k4

k5

k6

k7

k8

k9

0.05

2.2

11

100

0.11

900

300

500

300

Fig. 10. Proposed intermediate-pressure reaction network of 2-MTHF on Ni2P/SiO2 at 300 C, 0.5 MPa and 5% 2-MTHF and rate constants (Units in s1). Putative intermediates
are shown in brackets.

Conversion, Selectivity / %

80

2-MTHF onto the catalyst. Therefore, the conversion decreases as


H2 partial pressure increases, while the selectivity to n-pentane increases. This type of behavior is consistent with a LangmuirHinshelwood mechanism.
According to the LangmuirHinshelwood mechanism, the reaction may be described by the following equations.

n-Pentane

70
60
50
40
Conversion

30

n-Butane

20

2-Pentanone

10
0
0.0
2-Pentene

K MTHF

2-MTHF  2-MTHF
KH

2-Pentanol
1-Pentanol
0.5

1.0

1.5

2.0

H2 2 2H

2.5

Contact time / sec

Fig. 11. Analysis of proposed reaction. The points are the experimental results at
300 C, 0.5 MPa and 5% 2-MTHF, and the curves are the calculated ts.

single surface atoms are preferred. This results in an a-hydride


elimination from the carbon atom adjacent to the oxygen atom
to produce 2-pentanone. Similar intermediate steps can result
from ring-opening on the more hindered side (not shown). The results show that substantially different mechanisms can operate
depending on the conditions of reaction and the nature of the
surface.
3.2.6. Partial pressure analysis
Fig. 14 shows the conversions and the selectivity to n-pentane,
n-butane, 2-pentanone, 2-pentanol, and 1-pentanol. Fig. 14a-1 and
a-2 shows the results with varying H2 partial pressures, while
Fig. 14b-1 and b-2 shows the results with varying 2-MTHF partial
pressures. The raw data and analysis are in the Supplementary
information section, Part 5. From the H2 partial pressure analysis,
the conversion decreases as H2 partial pressure increases
(Fig. 14a-1). However, the selectivity to n-pentane increases, while
the selectivity to 2-pentanone decreases (Fig. 14a-2). In the case of
the 2-MTHF partial pressure analysis, the conversion decreases as
2-MTHF partial pressure increases (Fig. 14b-1), but the selectivity
to n-pentane also decreases (Fig. 14b-2). These results indicate that
H2 and 2-MTHF adsorbed competitively on the surface of the
catalysts. Higher H2 partial pressure is advantageous in hydrodeoxygenation but too much H2 interferes with the adsorption of

2-MTHF H ! P 2

In these equations, the symbol * indicates vacant sites on the catalyst surface, on which 2-MTHF and H2 are adsorbed, and 2-MTHF*
and H* represent adsorbed 2-MTHF and dissociatively-adsorbed
hydrogen atoms. Step (7) represents the adsorption of 2-MTHF, step
(8) represents dissociative adsorption of H2 on the catalyst surface,
and the Ki are adsorption equilibrium constants. Step (9) depicts the
rate-determining reaction of adsorbed 2-MTHF (2-MTHF*) and a
hydrogen atom (H*) on the catalyst surface to form a product. This
rate-determining step is the ring-opening reaction, which can occur
in two ways with surface hydrogen as shown earlier (Fig. 13). Here,
k is the rate constant of the surface reaction. Three assumptions are
made: (1) The surface reaction is the rate-controlling step, (2) The
reverse reaction can be ignored, and (3) The adsorption of products
can be ignored.
Two pieces of evidence give support for the assumption in the
LangmuirHinshelwood analysis that the adsorption of species
other than the reactants can be neglected. The rst comes from
the reaction network analysis (Fig. 11), which indicates that the
initial ring-opening is rate-determining. This indicates that subsequent steps, including desorption of the product species, are fast
and can be ignored in the kinetics. The second comes from the linear dependence of conversion on contact time (Fig. 9), which indicates that each reaction event of the 2-MTHF reactant in this range
of conversion is independent. This means that each reactant molecule sees a similar surface (mainly covered with hydrogen, with a
few empty sites), so that the conversion is not decreased by blockage of sites by adsorbed species.

25

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

SN2
k1
H

CH3

O+

Ni
O

CH3

CH3

Rate KhMTHF hH

Ni

CH3

O+

kK MTHF PMTHF h

Ni

k5

+
O

CH3
H

Ni

SN1

CH3

Ni

Fig. 12. Mechanism of Ring-Opening of 2-MTHF on Ni2P.

A rate equation as a function of partial pressure of 2-MTHF


and H2 was derived assuming that the coverages on the surface
are by 2-MTHF (hMTHF), hydrogen atoms (hH), and empty sites

TOFs1

25

Conversion

15
10
5
100

200

300

13

(b-1)
Conversion

15
10
5
0

10

20

30

40

(b-2)
70

60

60

50

2-Pentanone

30

n-Butane

20
10

Selectivity / %

Selectivity / %

p
kK MTHF PMTHF K H2 PH2
p
n
1 K H2 PH2 K MTHF PMTHF

400

n-Pentane

70

n-Pentane

50
40
30

2-Pentanone
n-Butane

20
10

2-Pentanol

2-Pentanol

1-Pentanol 100

12

20

(a-2)

40

11

Note that the exponent in the denominator was also set as a


variable and optimized. The results are shown in Table 3. Furthermore, the reaction rates (TOF) calculated with these optimized
values are shown as the curves in Fig. 15. The optimized value of

(a-1)

20

q
K H2 PH2 h

p
kK MTHF PMTHF K H2 PH2
p
2
1 K H2 PH2 K MTHF PMTHF

Conversion / %

25

10

The turnover frequency (TOF (s1)) is calculated from


TOF = (FA0  X)/(W  SCO) with 2-MTHF feed ow rate (FA0 (lmol/
s)), conversion (X), catalyst weight (W (g)) and the number of CO
uptake sites (SCO (lmol/g)). The TOF indicates the reaction rate
per active site, where the number of active sites are estimated by
the CO uptake. In Fig. 15, the plots show the TOF against the H2
partial pressure (a) and 2-MTHF partial pressure (b) calculated
from the experimental data. The TOF increases as the 2-MTHF
partial pressure increases, which is the opposite trend to conversion. This is because to attain higher 2-MTHF partial pressure,
the 2-MTHF feed ow rate was increased and this more than compensated for the decrease in conversion. The total ow rate was
kept constant in these measurements by adjusting the ow rate
of inert N2. Interestingly, selectivity to n-butane seems to not depend on the 2-MTHF partial pressure.
Using these experimental results, adsorption equilibrium constants (KMTHF and KH2) and the rate constant k were optimized
using POLYMATH 6.1. The equation used for optimization is shown
below.

Fig. 13. Proposed initial steps in the formation of primary intermediates. (a) Lowpressure route and (b) Medium-pressure route.

Conversion / %

(h*). The details are given in the Supplementary Information section, Part 6.

200

300

H2 partial pressure / kPa

400

10

20

30

40

1-Pentanol 2-MTHF partial pressure / kPa

Fig. 14. Conversions and selectivities to n-pentane, n-butane, 2-pentanone, 2-pentanol, and 1-pentanol at 300 C, 0.5 MPa (a-1) Conversions of H2 partial pressure analysis at
a 2-MTHF partial pressure of 9 kPa. (a-2) Selectivities of H2 partial pressure analysis. (b-1) and (b-2) those of 2-MTHF partial pressure analysis at a H2 partial pressure of
332 kPa.

26

A. Iino et al. / Journal of Catalysis 311 (2014) 1727

(a) 2-MTHF partial pressure : 9 kPa

(b) H2 partial pressure : 332 kPa


0.3

-1

0.2

TOF / s

TOF / s

-1

0.3

0.1

0.0
0

100

200

300

400

500

H2 Partial pressure / kPa

0.2

0.1

0.0
0

10

20

30

40

50

2-MTHF Partial pressure / kPa

Fig. 15. TOF from experimental results (plots) and calculation (lines) of (a) H2 partial pressure analysis and (b) 2-MTHF partial pressure analysis. Conditions: 300 C, 0.5 MPa
total pressure, balance N2.

Table 3
Fitting parameters and regression coefcient.
KMTHF (kPa1)
KH2 (kPa1)
k (s1)
n
R2

31
1.0  104
2.5
2.1
0.9

n was 2.1, which is close to 2, which is the theoretically derived value as shown in Eq. (13). This suggests that the rate-determining
step involves the reaction of adsorbed 2-MTHF with a single adsorbed hydrogen atom. This is consistent with the ring-opening
step shown in Fig. 12.
Furthermore, fairly good agreement between calculated and
experimental results can be seen. Comparing the two adsorption
equilibrium constants KMTHF and KH2, the H2 adsorption equilibrium constant KH2 is much larger than KMTHF. This indicates that
H2 is more strongly adsorbed on the catalyst surface than 2-MTHF.
Therefore, H2 will inhibit the rate and cause lower reaction rate
(TOF) and conversion. However, as shown in Fig. 14, selectivity to
oxygen-free products like n-pentane and n-butane increased as
H2 partial pressure was increased, and this means that after 2MTHF is adsorbed on the catalyst surface, higher H2 can carry
out the subsequent HDO reactions.
These ndings are also reected in the calculated coverages
(Supplementary Information, Part 5) which show the following order, hH > hMTHF  h*, that is, the hydrogen coverage is higher than
that of 2-MTHF with both much larger than the empty sites. The
2-MTHF thus sees mainly adsorbed hydrogen and a few empty
sites consistent with the overall picture for the surface presented
earlier.
4. Conclusions
A Ni2P/SiO2 catalyst was synthesized by temperatureprogrammed reduction of a phosphate precursor and was tested
in the hydrodeoxygenation (HDO) of 2-methyltetrahydrofuran (2MTHF) at an intermediate pressure of 0.5 MPa. Comparison of the
catalysts, Ni2P/SiO2, Pd/Al2O3 and Ru/C showed that Ni2P/SiO2
had higher selectivity to n-pentane and lower selectivity to lower
hydrocarbons with carbon numbers less than 5 than Pd/Al2O3
and Ru/C. This indicates that the Ni2P/SiO2 catalyst has high HDO
activity and low carbon bond cracking propensity. By contact time
analysis, the reaction network of HDO of 2-MTHF on Ni2P/SiO2 was
obtained. The rst steps are opening of the furanic ring by cleavage
of a CAO bond, and generation of 2-pentanone or pentanal. These
compounds which have a carbonyl group are converted into

pentanols by hydrogenation, pentenes by pentanol dehydrogenation, and then n-pentane by pentene hydrogenation. n-Butane is
generated as a coproduct with carbon monoxide by decarbonylation of pentanal. The model tting results agree well with the
experimental results. The magnitude of the kinetic parameters
indicated that the ring-opening step was the rate-controlling step
of this reaction. From the partial pressure analysis and theoretical
tting, 2-MTHF and H2 are shown to be competitively adsorbed on
the Ni2P/SiO2 catalyst with H2 more strongly adsorbed. Therefore,
H2 will cause lower conversion and reaction rate although higher
H2 partial pressure has an advantage in the subsequent deoxygenation reactions. These facts help to understand the nature of the
Ni2P/SiO2 catalyst and the HDO reaction mechanism.
Acknowledgments
This work was supported by Development of Next-generation
Technology for Strategic Utilization of Biomass Energy of New
Energy and Industrial Technology Development Organization
(NEDO), Japan and by the US Department of Energy, Ofce of Basic
Energy Sciences, through Grant DE-FG02-963414669.
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.jcat.2013.11.002.
References
[1] A.V. Bridgwater, Biomass Bioenergy 38 (2012) 68.
[2] P. Bui, J.A. Cecilia, S.T. Oyama, A. Takagaki, A. Infantes-Molina, H. Zhao, D. Li, E.
Rodrguez-Castelln, A. Jimnez Lpez, J. Catal. 294 (2012) 184.
[3] G. Yang, E.A. Pidko, E.J.M. Hensen, J. Catal. 295 (2012) 112.
[4] C. Zhao, S. Kasakov, J. He, J.A. Lercher, J. Catal. 296 (2012) 12.
[5] L. Ingram, D. Mohan, M. Bricka, P. Steele, D. Strobel, D. Crocker, B. Mitchell, J.
Mohammad, K. Cantrell, C.U. Pittman Jr., Energy Fuels 22 (2008) 614.
[6] C. Branca, P. Giudicianni, C.D. Blasi, Ind. Eng. Chem. Res. 42 (2003) 3190.
[7] R.H. Venderbosch, W. Prins, Biofuels Bioprod. Bioref. 4 (2010) 178.
[8] F. Agblevor, O.D. Mante, S.T. Oyama, R. McClung, Biomass Bioenergy 45 (2012)
130.
[9] O.D. Mante, F. Agblevor, S.T. Oyama, R. McClung, Appl. Catal. A: Gen. 445446
(2012) 312.
[10] P.M. Mortensen, J.-D. Grunwaldt, P.A. Jensen, K.G. Knudsen, A.D. Jensen, Appl.
Catal. A 407 (2011) 1.
[11] E. Laurent, B. Delmon, Appl. Catal. A 109 (1994) 77.
[12] S.B. Gevert, M. Eriksson, P. Eriksson, F.E. Massoth, Appl. Catal. A 117 (1994)
151.
[13] A. Centeno, E. Laurent, B. Delmon, J. Catal. 154 (1995) 288.
[14] A.Y. Bunch, U.S. Ozkan, J. Catal. 206 (2002) 177.
_ S
enol, E.-M. Ryymin, T.-R. Viljava, J. Mol. Catal. A 277 (2007) 107.
[15] O.I.
[16] A.L. Jongerius, R. Jastrzebski, P.C.A. Bruijnincx, B.M. Weckhuysen, J. Catal. 285
(2012) 315.
[17] A. Gutierrez, R.K. Kaila, M.L. Honkela, R. Slioor, A.O.I. Krause, Catal. Today 147
(2009) 239.

A. Iino et al. / Journal of Catalysis 311 (2014) 1727


[18] Y. Wang, Y. Fang, T. He, H. Hu, J. Wu, Catal. Commun. 12 (2011) 1201.
[19] C. Zhao, J. He, A.A. Lemonidou, X. Li, J.A. Lercher, J. Catal. 280 (2011) 8.
[20] L. Chen, Y. Zhu, H. Zheng, C. Zhang, B. Zhang, Y. Li, J. Chem. Technol. Biotechnol.
87 (2012) 112.
[21] X. Wang, P. Clark, S.T. Oyama, J. Catal. 208 (2002) 321.
[22] S.T. Oyama, J. Catal. 216 (2003) 343.
[23] R. Prins, M.E. Bussell, Catal. Lett. 142 (2012) 1413.
[24] V.M. Whiffen, K.J. Smith, Energy Fuels 24 (2010) 4728.
[25] S.T. Oyama, T. Gott, H. Zhao, Y.-K. Lee, Catal. Today 143 (2009) 94.
[26] K. Li, R. Wang, J. Chen, Energy Fuels 25 (2011) 854.
[27] H.Y. Zhao, D. Li, P. Bui, S.T. Oyama, Appl. Catal. A 391 (2011) 305.
[28] A. Cho, J. Shin, A. Takagaki, R. Kikuchi, S.T. Oyama, Top. Catal. 55 (2012) 969
980.
[29] Y.-K. Lee, S.T. Oyama, J. Catal. 239 (2006) 376.
[30] D. Li, P. Bui, H.Y. Zhao, S.T. Oyama, T. Dou, Z.H. Shen, J. Catal. 290 (2012) 1.
[31] S.T. Oyama, Y.-K. Lee, J. Catal. 258 (2008) 393.
[32] S.T. Oyama, X. Wang, Y.-K. Lee, K. Bando, F.G. Requejo, J. Catal. 210 (2002) 207.

27

[33] Y. Zao, M. Xue, M. Cao, J. Shen, Appl. Catal. B 104 (2011) 229.
[34] P. Grange, E. Laurent, R. Maggi, A. Centeno, B. Delmon, Catal. Today 29 (1996)
297.
[35] S.J. Sawhill, K.A. Layman, D.R. Van Wyk, M.H. Engelhard, C. Wang, M.E. Bussell,
J. Catal. 231 (2005) 300.
[36] J. Wildschut, F.H. Mahfud, R.H. Venderbosch, H.J. Heeres, Ind. Eng. Chem. Res.
48 (2009) 10324.
[37] D. Prochzkov, P. Zmostny, M. Bejblov, L. Cerveny, J. Cejka, Appl. Catal. A:
Gen. 332 (2007) 56.
[38] A. Oasmaa, E. Kuoppala, A. Ardiyanti, R.H. Venderbosch, H.J. Heeres, Energy
Fuels 24 (2010) 5264.
[39] R.H. Venderbosch, A.R. Ardiyanti, J. Wildschut, A. Oasmaa, H.J. Heeres, J. Chem.
Technol. Biotechnol. 85 (2010) 674.
[40] D.E. Resasco, J. Phys. Chem. Lett. 2 (2011) 2294.
erveny, J. C
ejka, Appl. Catal. A 296 (2005) 169.
[41] M. Bejblov, P. Zmostny, L. C
[42] T.T. Pham, L.L. Lobban, D.E. Resasco, R.G. Mallinson, J. Catal. 266 (2009) 9.

Вам также может понравиться