You are on page 1of 13

Fuel 117 (2014) 10611073

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Enhancement of phenol hydrodeoxygenation over Pd catalysts


supported on mixed HY zeolite and Al2O3. An approach to O-removal
from bio-oils
S. Echeandia a, B. Pawelec b,, V.L. Barrio a, P.L. Arias a,, J.F. Cambra a, C.V. Loricera b, J.L.G. Fierro b
School of Engineering (UPV/EHU), Chemical and Environmental Engineering Department, c/Alameda Urquijo s/n, 48013 Bilbao, Spain
Instituto de Catlisis y Petroleoqumica, CSIC, Cantoblanco, 28049 Madrid, Spain

g r a p h i c a l a b s t r a c t

 Phenol HDO was compared on Pd

100

l
A
YH
/2

0%

NiMo/
Al203-zeo
sulfided

Pd

80

HDO of phenol:
T =250 C
P =1.5 MPa
WHSV =0.5 h -1

H
0%

40

Y-

60

/1

/H

Pd
20

Pd

Phenol conversion (%)

catalysts supported on alumina,


zeolite HY and mixtures.
 The 20%HY80%Al2O3 support is
benecial for phenol HDO reaction
over Pd catalyst.
 The density of active phases on the
support surface inuenced on the
HDO activity.
 Catalyst acidity inuence on both
activity and coke formation.

h i g h l i g h t s

/A

Pd

Catalyst

a r t i c l e

i n f o

Article history:
Received 1 December 2011
Received in revised form 7 October 2013
Accepted 8 October 2013
Available online 23 October 2013
Keywords:
Phenol
HDO
HY zeolite
AluminaHY
Alumina

a b s t r a c t
This contribution describes the effect of the support (zeolite ultrastable HY, alumina (Al) and mixed
HYAl carriers) on the catalytic activity of Pd catalysts in the phenol hydrodeoxygenation (HDO) reaction
carried out in a ow xed-bed reactor at T = 523573623 K, P = 15 bar and WHSV = 0.5 h1. Phenol
dissolved in n-octane was used as model compound of bio-oil species derived from fast pyrolysis of lignocellulosic biomass. The catalysts were characterized by N2 physisorption, XRD, TPR, TPD-NH3, DRIFT
spectroscopy of adsorbed CO, HRTEM, X-ray photoelectron spectroscopy (XPS) and TPO/TGA techniques.
The largest phenol conversion (63%) achieved at 523 K over the reduced Pd/20%HYAl catalyst was similar to that obtained on a commercial NiMo/Al2O3zeolite hydrocracking sample (HCK) activated by sulfidation. Regardless of the reaction temperature, the only products detected in the HDO of phenol over all
catalysts studied were four O-free compounds: benzene, cyclohexene, cyclohexane, and methylcylopentene. Both reduced Pd/20%HYAl and sulded commercial HCK catalysts produced similar yields of O-free
products. From the catalyst activity-structure correlation, it can be concluded that the HDO of phenol is
favoured on the bifunctional Pd/20%HYAl catalyst which possesses moderate acidity and improved Pd
dispersion on the support surface. The contributions of the acid sites to the catalyst activity and deactivation by coke are discussed.
2013 Elsevier Ltd. All rights reserved.

Corresponding authors. Tel.: +34 915854769; fax: +34 915854760 (B. Pawelec),
Tel.: +34 946017282; fax: +34 946014179 (P.L. Arias).
E-mail addresses: bgarcia@icp.csic.es (B. Pawelec), pedroluis.arias@ehu.es (P.L.
Arias).
0016-2361/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.fuel.2013.10.011

1. Introduction
Mitigating greenhouse gases (GHGs) is one of the biggest challenges in the 21st century and requires long-term planning as well

1062

S. Echeandia et al. / Fuel 117 (2014) 10611073

as social awareness. Renewable sources of energy can contribute


signicantly to the effort of mitigating GHGs. Among different
renewable energy technologies, biomass-based technologies have
high potential [1]. Biomass has been proposed as a feedstock for liquid fuels and chemicals [2,3]. Many processes that transform biomass to liquid fuels start with the thermal breakdown of biomass
feedstock producing the so-called bio-oils [4]. The nal bio-oils
have higher energy density than the raw biomass precursor, which
makes it more suitable for industrial applications [5].
Pyrolysis bio-oils obtained from the fast pyrolysis of biomass
are complex mixtures of reactive chemical compounds (carboxylic
acids, aldehydes, ketones, carbohydrates, degraded lignin), water
(up to 25%), and some alkali and alkaline-earth metals (Na, K,
Mg, Ca) [6]. The high contents of oxygenated compounds (close to
40% oxygen) contribute to some deleterious properties, such as high
viscosity, corrosiveness, poor heating value, immiscibility with
hydrocarbon fuels, as well as to undesirable formation of carbon
deposits in parts of the automotive engines upon combustion, especially in compressionignition engines, i.e. diesel engines. In order to
improve the thermostability of these bio-oils, the oxygen removal
via hydrodeoxygenation (HDO), which is a similar process to hydrotreatment in a petroleum renery, is a common practice [7].
Past efforts for HDO catalyst development were focused on alumina-supported hydroprocessing catalysts, intensively used in
petroleum rening for several decades and therefore little attention has been paid to development of novel catalysts. Thus, the catalysts often studied for HDO were sulded CoMo/c-Al2O3 and
NiMo/c-Al2O3 systems [813]. Unfortunately, in the absence of a
sulding agent, the sulded catalysts deactivated quickly in HDO
reaction and the selectivity to different hydrocarbons changed
with time on-stream due to oxidation of the active phases [11].
Thus, great efforts have been devoted to enhance the stability of
these catalysts, but the results are not yet satisfactory [713]. Recent laboratory and commercial developments in the eld of catalytic hydroprocessing of biomass-derived liquefaction conversion
products are presented in a few excellent revisions [1419].
Recently, the use of noble metal catalysts as an alternative to
sulded catalysts for HDO reaction has been extensively studied
[1419 and references therein]. As pioneering patent demonstrated that Pd catalysts are highly effective for hydrogenation of
organic compounds typically found in bio-oils [20], more studies
were developed on Pd-based catalysts than in other noble metals
[1419 and references within]. Thus, the HDO activity of Pd-based
catalysts was investigated using acidic substrates such as ZrO2
[2123], Al2O3 [24,25], SiO2Al2O3 [25], SiO2 [2627], SBA-15 and
Al-SBA-15 [29], Al2(SiO3)3 [29], SAPO-31 [30], carbon [3139], zeolites Beta and ZSM-5 [37,38], MCM-41 [37], and super acid SO2
4 /
ZrO2/SBA-15 [40] substrates. On the contrary, studies of the HDO
reaction over Pd catalysts supported on basic carriers are scarce
[41,42]. The reason for this lies in the fact that the HDO reaction
over noble metals supported on acidic substrates requires both
metal site and acid sites, being H2 dissociated on metal sites while
O-containing compounds are adsorbed and activated on either metal sites or on the cations/oxygen vacancies located at the metalsupport interface [1419]. Moreover, it is well known that the
acidic sites of supports might to catalyse dehydration, isomerisation, alkylation and condensation reactions [43].
The effects of the substrate structure (C, H-Beta, ZSM-5, MCM41) and its acidity on activity of supported Pd catalysts in the
transformation of benzophenone and benzaldehyde were investigated by Cejka et al. [37,38] who found that the substrate structure
has a determining effect on the course of HDO reaction being the
most suitable one Pd catalysts supported on large pore zeolite Beta
[37,38]. This was conrmed also for Pt catalysts supported on HBeta [44] and HY [45,46] zeolites. Similarly, Pt catalysts supported
on mesoporous ZSM-5 showed better performance in dibenzofuran

HDO than its counterpart supported on microporous ZSM-5 zeolite


due to the diffusion limitation of dibenzofuran in the microporous
structure of ZSM-5 [44]. Considering the stability of substrate, a
particularly promising approach to the development of novel
HDO catalysts lies in the use of highly acidic ultrastable HY zeolite
substrate. This is because the ultrastable HY zeolite has the advantages of higher surface-to-volume ratio than alumina, variable
framework compositions and high hydrothermal stability [47].
Zeolite acidity, however, drastically leads to fast deactivation rates
and increases the amount of undesirable cracking products, which
accelerates the rate of coke deposition and the yield of gases [48].
In this sense, a large deactivation of bifunctional zeolite-supported
noble metal catalysts in the HDO of phenol was reported [45].
To the best of our knowledge, the only study on HY-supported
noble metal HDO catalysts was reported by Hong et al. [46] who
showed that bifunctional HY zeolite supported Pt catalysts are
effective for the aqueous phenol HDO using a xed-bed reactor
working under high H2 pressure (4 MPa). It was found that phenol
transformation over Pt/HY catalysts proceeds via hydrogenation
hydrogenolysis ring-coupling reactions producing monocyclic
and useful bicyclic hydrocarbons [46]. The bare HY zeolite demonstrated to be not effective catalyst for HDO of phenol upon reaction
conditions employed (T = 473523 K and PH2 = 4 MPa) conrming
that bifunctional catalysts are needed for HDO reaction: H+ sites
for dehydration and metal sites for hydrogenation are needed. Notwithstanding, it is emphasized that Pt/HY catalysts were activated
at high temperature (773 K) [46], thus sintering of Pt crystallites is
expected to occur on the surface of HY zeolite.
Taking into account that future commercial applications of lignin,
its conversion will require catalysts with improved resistance to
deactivation [44]. In the present work, we aimed to exploit the
advantage of the properties of the alumina and HY zeolite mixtures
as support for Pd in order to develop stable and more selective catalysts toward the hydrogen-donor species. Indeed, revision of the literature on the use of Pd catalysts for HDO reaction [2042] indicates
that hybrid HYAl substrate has not been used to prepare Pd HDO
catalysts. Thus, we compare here the activity for phenol HDO reaction of Pd catalysts supported on alumina, mixed alumina-HY zeolite
materials and alumina-free zeolite, trying to establish a relationship
between activity and catalyst structure. Moreover, considering that
the mechanism of hydrogenation over supported noble metal catalyst is still debated [1419], our selectivity results might contribute
to clarify this point. The physicochemical properties of the catalysts
have been evaluated by various techniques and their activity compared with those of a commercial zeolite-loaded alumina-supported
NiMo sulde hydrocracking catalyst to provide the catalyst with
hydrocracking function arising from substrate acidity.
2. Experimental
2.1. Catalyst preparation
Four supported Pd catalysts were prepared using c-Al2O3 (Girdler), HY zeolite (Conteka), and mixed 10%HY/c-Al2O3 as supports.
Two mixed HYAl2O3 supports were prepared by physical mixture
of different amounts (10 and 20 wt.%) of ultrastable HY zeolite
(Conteka) and c-Al2O3 (Girdler) in excess of distilled water under
vigorous stirring at room temperature for 2 h. The characteristics
of the HY zeolite are as follows: SiO2/Al2O3 mole ratio 5.6, Na2O
content 0.14 wt.% and unit cell 2.454 nm. After solid decantation,
the solid was dried at 383 K for 12 h and calcined at 523 K for
1 h. Depending on the composition, the supports will be abbreviated hereafter as (x)HYAl, where x is the nominal wt.% of HY
(10% and 20%). Before Pd incorporation, all substrates were dried
at 523 K for 1 h. Supported palladium catalysts were prepared by
wet impregnation of the respective substrate with aqueous

1063

S. Echeandia et al. / Fuel 117 (2014) 10611073

solutions of H2PdCl6 (Sigma) with the appropriate concentration to


achieve a metal content of 1 wt.%. A known weight of carrier was
contacted with a solution containing the required amount of palladium added to 100 mL of water, and the pH was xed at a value
close to 7. After the adsorption equilibrium was reached (contacting for 12 h at room temperature), the excess of water was removed in a rotary evaporator till dryness. Subsequently, the
impregnate was dried at 383 K in air for 12 h, and nally calcined
at 573 K for 4 h. No chloride was detected by XPS and EDX/TEM
techniques.
2.2. Catalyst characterization
2.2.1. Scanning electronic microscopy
The morphology of all calcined catalysts was studied by scanning electronic microscopy (SEM) on a Hitachi Tabletop TM-1000
electron microscope, equipped with energy dispersive X-ray analysis EDX QUANTAX 50.
2.2.2. Wide angle X-ray diffraction (XRD)
Calcined catalysts were characterized by powder X-ray diffractometry according to the step-scanning procedure (step size 0.02;
0.5 s) with a computerized Seifert 3000 diffractometer, using Ni-ltered Cu Ka (k = 0.15406 nm) radiation and a PW 2200 BraggBrentano h/2h goniometer equipped with a bent graphite monochromator
and an automatic slit. The assignment of the various crystalline
phases was based on the JPDS powder diffraction le cards.
2.2.3. N2 adsorptiondesorption isotherms
The textural properties of the oxide catalysts and bare supports
were determined from the adsorptiondesorption isotherms of
nitrogen recorded at 77 K with a Micromeritics TriStar 3000 apparatus. Prior to measurements, the samples were degassed at 573 K
under vacuum for 5 h. Specic surface areas of the supports and
supported Pd catalysts were calculated by the BET method from
the N2 adsorptiondesorption isotherms. The normalized SBET
(NSBET) was calculated from Eq. (1):

NSBET SBET ox =1  ySBET sup 

where SBET ox and SBET sup are the values of BET areas of the oxide
catalyst and support, respectively, and y is the Pd loading.
2.2.4. Chemical analyses
The metal loading of the calcined catalysts was determined by
Inductively Coupled Plasma Atomic Emission Spectroscopy (ICPAES), Perkin Elmer Optima 3300DV. The solid samples were rst
digested (in a mixture of HF, HCl and HNO3) in a microwave oven
for 2 h. Then, aliquots of solution were diluted to 50 mL using
deionized water (18.2 mX quality). The metal loadings of calcined
catalysts are given in Table 1.
2.2.5. Temperature-programmed techniques
Temperature-programmed reduction (TPR) was performed on
the oxide precursors using a semiautomatic Micromeritics TPD/

TPR 2900 apparatus interfaced to a microcomputer. Prior to reduction, the catalyst (0.05 g) was dried in a TPR cell at 573 K for 1 h in
a stream of helium to remove water and other contaminants. TPR
proles were obtained by passing a 10% H2/Ar ow (50 mL min1)
through the sample. The temperature was increased from room
temperature to 1273 K at a rate of 15 K min1, and the amount of
H2 consumed was determined with a thermoconductivity detector
(TCD). The efuent gas was passed through a cold trap before the
TCD in order to remove water from the exit stream.
The acidity of the fresh reduced catalysts was determined by
temperature-programmed desorption (TPD) of ammonia using
the same equipment employed for TPR experiments. A sample
(0.05 g) was dried in a TPD cell at 573 K for 1 h in a stream of helium, reduced with H2 at the same temperature for 1 h, cooled
down in He ow to 373 K, and then exposed to ammonia for
0.5 h. TPD measurements were started from 373 K at a heating rate
of 15 K min1 till 873 K using helium as a carrier gas
(50 mL min1). In order to determine the total acidity of the catalyst from its NH3 desorption prole, the area under the curve
was integrated. A semiquantitative comparison of the strength distribution was archived by Gaussian deconvolution of the peaks.
Weak, medium and strong acidities were dened as the areas under the peaks at the lowest, medium and highest temperatures,
respectively.

2.2.6. HRTEM measurements


The reduced (573 K, H2) catalysts were studied by HRTEM
microscopy using a JEM 2100F microscope operating with a
200 kV accelerating voltage and tted with an INCA X-sight (Oxford Instruments). Energy dispersive X-ray microanalysis (EDX)
system was used to verify the semi-quantitative composition of
the supported phases. The catalysts were ground into a ne powder and dispersed ultrasonically in hexane at room temperature.
Then, a drop of the suspension was put on a lacey carbon-coated
Cu grid. In order to obtain statistically reliable information, the
particle size distribution was evaluated from several micrographs
taken from the same sample. The average particle size was estimated using the equation d = Rnidi/Rni, where ni is the number
of particles with diameter di and Rni is the number of particles
used to build the size distribution.

2.2.7. X-ray photoelectron spectroscopy (XPS)


The chemical state and surface composition of fresh reduced
(573 K, H2) catalysts were revealed by X-ray photoelectron spectroscopy (XPS). XP spectra were recorded with a VG Escalab 200R
spectrometer equipped with Mg Ka X-ray source (hm = 1254.6 eV).
The procedure followed during XPS measurements was described
elsewhere [48]. The binding energies were calculated with respect
to the C(C,H) component of the C1s peak xed at 284.8 eV. The
intensity of Pd 3d peaks was estimated by calculating the integral
of each peak after subtraction of an S-shaped background and tting the experimental peak to a Lorentzian/Gaussian lines (90G/
10L).

Table 1
Palladium contenta and textural propertiesb,c of pure supports and oxide precursors.

Catalyst

Pda (wt%)

SBETb (m2 g1)

NSBETc

Vtotal (cm3 g1)

dc (nm)

Pd/Al
Pd/10%HYAl
Pd/20%HYAl
Pd/HY

0.81
0.77
0.80
0.77

154
179
194
476

0.96
0.99
0.91
0.81

0.49
0.41
0.37
0.35

12.7 (7.5)
11.0 (11.4)
10.2 (12.0)
8.9 (7.1)

(163)
(185)
(216)
(594)

(0.33)
(0.44)
(0.46)
(0.29)

As determined by ICP-AES.
As determined by N2 adsorptiondesorption isotherms at 77 K (values in parentheses correspond to bare support); SBET: specic BET surface area; Vtotal: total pore
volume, d: average pore diameter.
c
NSBET: normalized SBET calculated from equation NSBET = (SBET of oxide catalyst/[(1  y)  SBET of support], where y is the Pd loading determined by ICP-AES.
b

1064

S. Echeandia et al. / Fuel 117 (2014) 10611073

2.2.8. Coke quantication


The coke deposited on the spent catalysts was quantied by
thermogravimetry (Mettler Toledo TGA/SDTA851 equipment) by
comparison of the TPO/TG proles of spent catalysts with those
of their fresh counterparts. The TPO experiments were carried
out by raising sample temperature up to a nal temperature of
1173 K at a rate of 10 K min1 in a 20% O2/N2 gas mixture
(50 mL min1).
2.3. Catalytic activity testing
Activity tests were performed using 0.2 g catalyst with particle
size 0.250.4 mm. An industrial NiMo hydrocracking (HCK) catalyst was used as reference. Prior to activity tests, catalysts were
activated by in situ reduction under a 1:3 H2:N2 gas mixture with
a hydrogen ow of 2.5 L h1 (NTP), and at 1 bar total pressure.
The reference NiMo/Al2O3zeolite catalyst was sulded with
10%H2S/90%H2 gas mixture at 673 K for 4 h. The catalysts were
heated at a rate of 10 K min1 from 298 K to 673 K, held at this
temperature for 4 h under the reduction gas mixture, and then
purged in owing N2 for 1 h. Activity measurements were carried
out in a bench-scale xed-bed catalytic reactor (9 mm i.d. and
300 mm length) placed inside a programmable furnace (Microactivity-Reference). A K-type thermocouple was used to measure
the temperature of the catalyst bed. Two mass ow controllers
maintained the hydrogen and nitrogen (diluent) ows. Prior to
each run, the system was kept under an inert atmosphere (N2),
while the pressure was increased up to 15 bar. The catalytic bed
was heated to the desired temperature and the hydrogen and the
liquid feed mixture were fed to the reactor. The experimental conditions were as follows: T = 523573623 K, total pressure 15 bar,
weight hourly space velocity (WHSV) of 0.5 gphe(gcat h)1 and a
feed consisting of 1 wt.% phenol (Merck > 99%), in n-octane (Fluka
Chemika > 99%) used as solvent was introduced at a ow of
10 g h1 with a H2:phenol molar ratio of 100:1. n-octane was used
as solvent because blank experiments proved that, contrary to
other solvents, this solvent did not react under the reaction conditions employed. Moreover, considering its non-polar character, a
lower interaction of n-octane with acid sites of support, as compared with more polar hydrophilic solvent, is expected [38].
The steady-state was usually achieved within 4 h on stream,
and the representative composition of the reaction products was
calculated from several subsequent analyses. The on-line analysis
of the efuent from the reactor was performed using a gas chromatograph (HP 5890, capillary column DB-1) equipped with a
ame ionization detector. For product identication a gas chromatograph (Agilent 5973, capillary column HP-5 MS) equipped
with a mass selective detector was used. The main detected products of phenol HDO were benzene, cyclohexane, cyclohexene and
methylcylopentane. The steady-state HDO activities were
expressed by overall phenol conversion (Xphe), where Xphe was
dened as 1  nphe =n0phe and n0phe and nphe are the number of moles
of phenol in the feed and the reaction products, respectively.
Duplicate experiments showed Xphe values within 3% deviation.
The carbon balances were calculated on the basis of liquid and
gas product analyses. The closures obtained were about 95% on
average.
3. Results
3.1. Characterization of oxide precursors
3.1.1. Physicochemical characteristics
The morphology of four Pd oxide precursors was studied by
SEM technique. The SEM images of those samples are shown in

Fig. S1 in the Supporting information (SI). SEM observations of


the Pd/10%HYAl and Pd/20%HYAl samples revealed non-uniform
distribution of HY zeolite on the alumina being Pd deposited preferentially on the HY surface rather than on alumina, as conrmed
by EDX/SEM elemental analysis (data not shown here). Pd/HY sample showed more uniform and smaller grains than its Pd/Al
counterpart.
Table 1 summarizes the main physicochemical characteristics
of all oxide precursors and bare supports. Regardless of the support, all oxide precursors showed similar palladium loading
(0.770.81 wt.%) which is close to the nominal one (1.0 wt.%).
The textural properties of oxide precursors were evaluated from
nitrogen adsorptiondesorption isotherms. Fig. 1(a) and (b) shows
the N2 adsorptiondesorption isotherms of the bare supports and
oxide precursors, respectively. Bare HY and Pd/HY samples exhibited type I isotherm with a hysteresis loop which, according to the
IUPAC classication, belong to type H4 [49]. The almost horizontal
plateau of Pd/HY isotherm is characteristic of an ideal microporous
solid, although some mesoporosity could be also observed. The latter is probably created during the hydrothermal treatment of the
HY zeolite. Contrary to the Pd/HY, the N2 adsorptiondesorption
isotherms of the Pd catalysts supported on both mixed HYAl2O3
substrates are of type IV with H1 hysteresis loop. This means that
both catalysts are mostly mesoporous. The hysteresis loop H1 has a
narrow loop with two branches almost parallel and vertical. This
type of hysteresis is usually found on solids consisting of agglomerates or aggregates of particles having a narrow pore size distribution. The N2 adsorptiondesorption isotherm of Pd/Al belong to
type IV with H2-type hysteresis loop, characteristics of solids with
ink-bottle pores [50].
The values of BET area, total pore volume and the average pore
diameter are also summarized in Table 1. As it can be seen, the BET
surface area follows the trend: Pd/HY  Pd/20%HYAl > Pd/
10%HYAl > Pd/Al whereas for the total pore volume the trend is
different: Pd/Al > Pd/10%HYAl > Pd/20%HYAl > Pd/HY. The highest BET area and the lowest total pore volume of the Pd/HY catalyst
is due to the large microporous region of HY zeolite (Fig. 1(a)). As a
consequence, this catalyst showed 3-fold higher BET surface area
than its Pd/Al counterpart (476 vs. 154 m2 g1). The increase in
BET surface area when increasing the HY zeolite content in mixed
HYAl2O3 materials is due to a larger contribution of the HY zeolite
microporosity system. Considering the relatively low zeolite content in the composite carriers, a relatively small increase in N2
sorption capacity upon raising HY content was observed (Table 1).
As expected, impregnation of the carriers with H2PdCl6 salt precursor leads to a decrease in BET area of the catalysts as compared to
the bare support indicating some pore blocking by palladium species. Additionally, one might expect that some dissolution of the
support occurs during impregnation of carrier with an acidic aqueous solution of H2PdCl6, similar to what was observed previously
for the preparation of the RuCl3/Al2O3 catalysts [51,52].
In order to ascertain the pore modication induced by the addition of HY zeolite to alumina, the BarretJoynerHalenda (BJH) formula was employed to determine pore size distribution (PSDs) of
the oxide precursors in both meso- (250 nm) and macroporous
(>50 nm) regions. Unfortunately, the BJH method is imprecise in
the region below 1 nm, where the supercages of the HY zeolite appear. The BJH pore size distribution of the oxide precursors is
shown in Fig. 2. As expected, the pore size distribution changed
going from Pd/Al to Pd/HY sample. The Pd/Al shows a very broad
pore size distribution between 2 and 70 nm, with a maximum at
17 nm, however the PSD of Pd/HY catalyst is narrower (230 nm)
and the maximum peaks at ca. 10.9 nm. For the Pd/10%HYAl
and Pd/20%HYAl catalysts, PSD showed a bimodal distribution
with maxima peaking at about 14.3 and 13.7 nm, respectively indicating that the average pore diameter decreased with an increase

1065

S. Echeandia et al. / Fuel 117 (2014) 10611073

350

(a)

Volume adsorbed (cm /g STP)

Pd/Al
Pd/10%HY-Al
Pd/20%HY-Al
Pd/HY

250

Volume adsorbed (cm /g STP)

Al
10%HY-Al
20%HY-Al
HY

250

(b)

300

type IV- H1 hysteresis

300

200
type I - H4 hysteresis

150

100

50

200
0.41

150

100

50
0

20HY-Al: type IV/ H2 hysteresis

0.0

0.2

0.4

0.6

0.8

1.0

Relative pressure P/P0

0
0.0

0.2

0.4

0.6

0.8

1.0

Relative pressure P/P0

Fig. 1. N2 adsorption-desorption isotherms at 77 K of bare supports (a) and oxide catalyst precursors (b).

dV/ dlog(D) Pore volume (cm /g/nm)

of HY content in mixed HYAl2O3 substrates. Considering the average pore diameter compiled in Table 1, the pore diameter follows
the trend: Pd/Al (12.7 nm) > Pd/10%HYAl (11.0 nm) > Pd/20%HY
Al (10.2 nm) > Pd/HY (8.9 nm). This trend clearly indicates that
loading HY zeolite into alumina leads to a decrease of the average
pore diameter, as it could be expected.
To evaluate the inuence of Pd loading on the location of Pd species within the support structure, the normalized SBET (NSBET) was
calculated employing Eq. (1). For the Pd/Al and both Pd catalysts
supported on mixed HYAl2O3 materials, NSBET values in 0.910.99
range suggest the presence of Pd species on the support surface.
For the Pd/HY sample, the NSBET value of 0.8 suggests the presence
of Pd species on the support surface as well as their location within
the inner supports porous structure (probably in the supercages of
the HY zeolite). Thus, the comparison of the normalized SBET values
indicates that, regardless of the support, most of the palladium species are accessible for reactant molecule (phenol).

3.1.2. Wide-angle XRD


X-ray powder diffraction data of the calcined catalysts were
used to identify crystalline palladium species formed on the support surfaces (Fig. 3). Regardless of the support composition, no
peaks of PdO or Pd0 crystallites were observed indicating their
amorphous character or that crystallite sizes are below the detection limit of the XRD technique (<4 nm). The X-ray diffraction patterns of the Pd/Al and Pd/HY catalysts were typical of the Al2O3
(JCPDS 00-001-1303) and faujasite (JCPDS 00-0111-0672) phases,
respectively. As expected, both Pd/10%HYAl and Pd/20%HYAl
catalysts showed both phases simultaneously. Thus, all Al-containing catalysts exhibited typical diffraction lines of Al2O3 material at
2h values of 37.5, 39.4, 45.8, 60.8, 66.9 and 85.0 whereas the
HY-containing catalysts showed many narrow peaks at 2h values of
6.1, 10.3, 12.2, 15.6, 18.4, 20.3, 23.7, 27.0, 31.8, 34.2, 38.4
and 54.6, which are typical of zeolite faujasite. As expected, comparison of XRD proles of the zeolite peaks in the HY-containing

Pd/HY
Pd/10%HY-Al
Pd/20%HY-Al
Pd/Al

17.0 nm

o o

o
o

Al2O3
HY

Linear counts (au)

o
14.3 nm
13.7 nm

o
Pd/HY

45.8

66.9

39.4

*
*

37.5

o o oo

60.8

85.0

o
o o o

Pd/20%HY-Al

10.9 nm

Pd/10%HY-Al

Pd/Al

20

40

60

80

100

120

Pore diameter (nm)

20

40

60

80

100

120

2 Theta ()
Fig. 2. BJH pore size distribution of Pd-based oxide precursors (from adsorption
branch) calculated by the Harkins and Jura equation.

Fig. 3. XRD diffraction patterns of Pd-based oxide precursors.

1066

S. Echeandia et al. / Fuel 117 (2014) 10611073

catalysts revealed an increase of their intensity with an increase of


HY content in the mixed HYAl2O3 supports.

Table 2
Hydrogen consumption during TPR of calcined Pd catalysts.
Catalyst

3.1.3. Temperature-programmed reduction


TPR proles of bare supports and oxide precursors are shown in
Fig. 4. As expected, contrary to bare supports, Pd catalysts showed
H2-consumption in the temperature range 273573 K. Thus, the
H2-consumption peak at 289 K observed for Pd/HY catalysts is
due to reduction of highly dispersed PdOx species [53,54] whereas
the negative H2 consumption peak at 358 K detected on Pd/Al and
the two Pd(x)HYAl catalysts is usually ascribed to the desorption
of hydrogen from a bulk palladium hydride formed through hydrogen diffusion into Pd crystallites [55]. This behaviour is typical for
catalysts with low metal dispersion due to the high palladium concentration in the large particles. The quantication of H2 consumption/evolution is listed in Table 2. From these data, the amount of
hydrogen desorbed from bulk palladium hydride follows the trend:
Pd/10%HYAl  Pd/Al > Pd/20%HYAl  Pd/HY (none). The absence of bulk b-PdH0.6 species on the surface of Pd/HY catalyst is
in agreement with TEM data which indicate the formation of smaller size Pd0 particles on this sample.
The Pd catalysts exhibited two types of reduction peaks: one at
low-temperature, which is due to reduction of easy reducible palladium oxide, and the other broad one located at somewhat higher
temperature (473573 K) due to reduction of a more stable PdO
species interacting strongly with the support [56,57]. For the Pd/
Al and Pd/10%HYAl catalysts, the peak maxima appeared around
501 and 511 K respectively, indicating similar crystallite sites. For
the Pd/20%HYAl and Pd/HY catalysts, this peak is very broad
and shifted to lower temperature when compared with Pd/Al and
Pd/10%HYAl catalysts. Indeed, considering that reduced Pd/HY
samples showed the lowest Pd0 particle among the catalysts studied (HRTEM results vide infra) and that bulk Pd-hydride is not
formed, the shift of its peak maxima toward lower temperature
424
291

(d)

x 10

Pd/HY

Amount of H2 (mmol/gcat)

Pd/Al
Pd/10%HYAl
Pd/20%HYAl
Pd/HY

289 K

358 K

473573 K

Total

0.89

0.24
0.26
0.21

1.68
1.82
1.96
2.44

1.44
1.56
1.75
3.33

can be explained considering that PdO clusters are rstly reduced


to form Pd atoms at lower temperature (H2 consumption at
291 K) and then H2 is dissociated over these Pd atoms and then
the H-atoms spilt over on the surface reduce the highly dispersed
Pd2+ species (H2 consumption at 424 K). Finally, an enhanced
reduction of Pd species on the Pd/HY can be explained considering
the H2 dissociation on reduced palladium particles and subsequent
hydrogen migration to the carrier interface (spill-over effect).
3.2. Characterization of fresh reduced catalysts
3.2.1. TPD-NH3
In order to compare the acidity of samples before and after Pd
incorporation, in Fig. 5 the TPD-NH3 proles of the reduced Pd catalysts are compared with that of the corresponding bare supports
pre-treated at the same conditions (H2/Ar reduction at 573 K for
1 h). These proles were mathematically tted using Gaussian
functions (not shown here). Based on the desorption temperature,
acid sites were identied to possess weak (T < 523 K) and medium
strength (523773 K). Regardless of support, all bare supports and
calcined Pd catalysts show weak and medium strength acid sites
which could be identied as Brnsted and structural type, respectively [58]. The concentration of acid sites (expressed as
mmolNH3 =gcat ) is shown in Table 2. From these data, it is seen that
Pd loading on acidic supports results in the loss of an important
fraction of acid sites. This is due to the loss of OH groups upon
Pd2+ ions attachment to the support surface. The total acidity of
the catalysts follows the trend: Pd/HY > Pd/20%HYAl > Pd/
10%HYAl > Pd/Al. This trend clearly indicates that the incorporation of HY zeolite into alumina carrier led to an increase acidity.

444
358

x 10

Pd/
20%HY-Al

511

(d)

x 10

(b)
497

Pd/
10%HY-Al

TCD signal (a.u.)

H2 consumption (a.u.)

(c)

(c)

(b)

x 10

(a)

Pd/Al
(a)

200

300

400

500

600

Temperature (K)
Fig. 4. TPR proles of bare supports (doted line) and Pdbased oxide precursors
(solid line) Pd/Al, (b) Pd/10%HY-Al, (c) Pd/20%HY-Al, (d) Pd/HY.

100

200

300

400

500

600

Temperature (C)
Fig. 5. TPD-NH3 proles of bare supports (doted line) and reduced Pd-based
catalysts (solid line): (a) Pd/Al, (b) Pd/10%HY-Al, (c) Pd/20%HY-Al, (d) Pd/HY.

S. Echeandia et al. / Fuel 117 (2014) 10611073

3.2.2. HRTEM
HRTEM was employed to reveal Pd distribution on the different
supports. The HRTEM images of the fresh reduced catalysts are displayed in Fig. 6. Examination of HRTEM images show that, in general, Pd crystallites are heterogeneously distributed on the support
surface but they become more homogeneous upon increasing zeolite content. The surface density of the Pd0 particles, expressed as
Pd0 particles per 50 nm2, is presented in Table 4. As seen in this table, the surface density of the Pd0 particles follows the trend: Pd/
HY  Pd/Al > Pd/20%HYAl > Pd/10%HYAl (Table 4). Unfortunately, the poor contrast between HY and alumina substrates does
not allow to unambiguously discriminate the preferential location
of palladium particles on either HY or alumina phase. In this sense,
the EDX/SEM elemental analysis gave information that Pd is deposited preferentially on the HY surface rather than on alumina (data
not shown).
The inuence of support on the Pd crystallite size distribution
was determined by statistical analysis of various HRTEM images
(see Fig. S2 in the Supporting Information). For the all catalysts,
the Pd particle sizes are distributed over the range of 210 nm
yielding the average Pd size in the range 3.25.7 nm (Table 4).
For the Pd particle size, the observed trend is: Pd/HY
(3.2 0.7 nm) < Pd/20%HYAl
(4.6 1.2 nm) < Pd/10%HYAl
(5.0 1.7 nm) < Pd/Al (5.7 2.2 nm). Thus, in good agreement with
TPR results (Fig. 4), the Pd/HY catalyst shows the lowest Pd particle
size among the catalysts studied. It is emphasized that information
on the Pd particle sizes obtained by TEM is different from XRD results (Fig. 3). This is due to difference in the detection limit and
associated error to XRD and TEM measurements. The XRD technique is not able detect regular crystal particles with size lower
than 4 nm whereas the TEM technique can do it. In addition,

Pd/Al

Pd/20%HY-Al

1067

TEM technique does not allow distinguish very small particles


(<1 nm) from the zeolite background because of the lack of contrast. It is also emphasized that the catalyst area irradiated by
the electron beam is several orders of magnitude than in case of
the XRD analysis, therefore the Pd size calculated by statistical
analysis may not be representative for the whole sample. We can
say that the number of small Pd clusters is underestimated by
TEM and that this leads to an overestimation of the average Pd
crystallite size.
3.2.3. XPS and DRIFT spectroscopy of adsorbed CO
Information on the Pd state and the surface composition of the
reduced (573 K, H2) catalysts was obtained by XPS analysis. Fig. 7
shows the Pd 3d core-level spectra of all reduced catalysts and Table 4 summarizes the Pd 3d5/2 binding energies (BEs). For all zeolite-containing catalysts, the binding energy of Pd 3d5/2 core level
at 334.9 eV (335.0 eV for Pd/HY) is due to metallic Pd [58,59].
The BE of the Pd3d5/2 peak in Pd/Al catalyst was 0.5 eV lower than
that measured for Pd/10%HYAl and Pd/20%HYAl catalysts. This
BE shift of Pd might involve Pd particle size effect. Considering
the TPR results, the possible contribution of the metal-support
interaction to the BE shift can be precluded.
CO adsorption followed by DRIFT spectroscopy was performed
in order to identify the nature of surface sites on the supported
Pd catalysts after their activation in pure hydrogen at 573 K. The
DRIFT-CO spectra (irreversible CO adsorption at room temperature) are displayed in Fig. S3 in the Supporting Information. As
compared with Al-containing samples, the Pd/HY catalyst showed
signicant differences in: (i) the frequency of the observed maxima, (ii) the total integrated intensity of the bands, and (iii) the ratio between different components. For the pre-reduced Pd/HY, the

Pd/10%HY-Al

Pd/HY

Fig. 6. HRTEM of the pre-reduced Pd catalysts.

S. Echeandia et al. / Fuel 117 (2014) 10611073

3.3. Activity measurements


Fig. 8(a) shows the activities of the reduced Pd/Al, Pd/10%HY
Al, Pd/20%HYAl and Pd/HY catalysts in the HDO of phenol at total
hydrogen pressure of 15 bar and three different temperatures (523,
573 and 623 K). The activity of a commercial NiMo sulde catalyst
used as reference is also included in this gure. Conversion data are
the average of the conversion achieved along 6 h of time on-stream
(TOS) reaction at each temperature (during this time the conversion changes were <4%). It was found that the optimal reaction
temperature was 523 K. This may be related to a faster deactivation by coking at higher temperatures. Because a low reaction tem-

Pd

Pd3d
3d5/2

counts per second (au)

3d3/2

Pd/Al

Pd/10%HY-Al

Pd/20%HY-Al

Pd/HY

330

335

340

345

BE (eV)
Fig. 7. Pd 3d3/2 core-level spectra of in-situ pre-reduced (300 C, H2) Pd catalysts.

100

(a)
Phenol conversion (%)

DRIFT spectra of adsorbed CO suggest that palladium would be in


an electron-decient state. However, considering TPR results, the
electronic effects associated with a strong metal-support interaction are hardly possible. Thus, the upward shift of a band indicative
of linear CO adsorption on Pd/HY zeolite may be related rather to
changes in the Pd particle size. Interestingly, both Pd/HY and Pd/
20%HYAl showed a largest contribution from CO linearly bonded
to metallic Pd species. For the Pd/Al, Pd/10%HYAl and Pd/20%HY
Al catalysts, the lowering of 34 cm1 of the CO-bridge frequency
with respect to Pd/HY (see Fig. S3 in the Supporting information
(SI)) probably reects the variation in the coordination of Pd atoms
in terraces, steps or kink sites.
Quantitative XPS data of all catalysts are summarized in Table 4.
Surface exposure of Pd species as evaluated by the Pd/(Si + Al)
atomic ratio follows the trend: Pd/Al  Pd/10%HYAl < Pd/
20%HYAl < Pd/HY. Therefore, both Pd/HY and Pd/20%HYAl catalysts exhibit a lower palladium surface exposure than Pd/Al and
Pd/10%HYAl ones. This difference can be also related to a higher
amount of Pd species located within the pore structure of the HY
zeolite. An additional evidence for this hypothesis comes from
the linear correlation found between palladium surface exposure,
determined by XPS, and the location of palladium species on the
support surface, derived from the normalized BET area of catalysts
(see Fig. S4 in the Supporting information (SI)).

Pd/Al
Pd/10HY-Al
Pd/20HY-Al
Pd/HY
sulfided NiMo

Reference

80

60

40

20

0
523 K

573 K

623 K

Temperature (K)

Yields of products at 523 K

1068

0,6

(b)

Benzene
Cyclohexene+cyclohexane
Methylcyclopentane

0,5

0,2

0,0

Pd/Al

Pd/10%HY-Al

Pd/20%HY-Al

Pd/HY

NiMo

Catalyst
Fig. 8. (a) Comparison of the steady-state phenol conversions at different temperatures and yields of products (b) in HDO of phenol at 250 C over supported Pd
catalysts and NiMo reference sample (P = 1.5 MPa, WHSV = 0.5 h1).

perature favours hydrogenation of the aromatic ring of Ocontaining compounds [25] leading to preferential phenol transformation via hydrogenation reaction route, the catalyst activity
was not studied at temperatures lower than 523 K. Indeed, only a
small amount of benzene was formed in the HDO of phenol at
473 K and 4 MPa over a Pt/HY catalyst [46].
Regardless of the reaction temperature, the highest phenol conversion was achieved for the Pd catalyst supported on the 20% HY/
Al2O3 catalyst. Incorporation of a smaller amount of HY (10 wt.%)
into alumina generated less active catalysts, and when pure HY
zeolite was used as support activity was the lowest with the exception of the test at 623 K. At 523 K, the Pd/20%HYAl catalyst
showed similar phenol conversion than the sulded NiMo/Al2O3
zeolite hydrocracking catalyst (63% vs. 62%) selected as reference
(HCK).
Table 5 shows the selectivity obtained on the different Pd catalysts in HDO of phenol reaction at 523, 573 and 623 K. Regardless
of the reaction temperature, the only products detected in the HDO
of phenol over all palladium catalysts were four O-free compounds: benzene, cyclohexene, cyclohexane and methylcyclopentane. Based on the selectivity data shown in Table 5, the possible
HDO reaction pathways for phenol HDO on Pd-based catalysts is
shown in Scheme 1. The proposed reaction scheme is based not
only in product distributions but also in literature information
[1419 and references within]. For all catalysts, the HDO of phenol
might occur via two parallel pathways: a hydrogenation (HYD) of
phenols aromatic ring followed by cleavage of the CO r bond
leading to the formation of cyclohexane and a direct cleavage of
the CO r bond leading to the formation of benzene [1419]. This
reaction scheme suggests that the use of hydrogen for phenol
transformation over HY-containing Pd has two effects with respect

S. Echeandia et al. / Fuel 117 (2014) 10611073

1069

Scheme 1. Scheme of the phenol HDO reaction over Pd/Al, Pd/10%HYAl, Pd/20%HYAl and Pd/HY catalysts.

to reaction mechanism: saturating double bonds and removing


oxygen. Because cyclohexanol was not detected on all catalysts
studied, in the reaction Scheme 1 this product is considered as
an intermediate which transforms into cyclohexene and cyclohexane. Since all catalysts showed 100% selectivity toward O-free
products, the hydrogen consumption for saturation of double
bonds could be deduced from selectivity toward cyclohexane formation. In the HDO reaction at 523 K, selectivity toward cyclohexane follows the trend: Pd/10%HYAl (97%)  Pd/20%HYAl
(95%)  Pd/Al (70%)  Pd/HY (55%) indicating that both Pd catalysts supported on hybrid substrates show larger hydrogen consumption for saturation of double bonds. While testing the Pd/
HY catalyst, relatively large quantity of methylcyclopentane (only
at 573 and 623 K) was detected. This product could be originated
as a result of combined effect of the high hydrogenation/dehydrogenation activity of the Pd and the acidity of the support.
Data collected in Table 5 indicates that there is a strong effect of
support (Al2O3, mixed HYAl2O3 and HY zeolite) on product selectivity. Regardless of the reaction temperature, product distribution
observed for the Pd catalysts supported on mixed supports
(10 wt.% and 20 wt.% of HY) yielded a high proportions of cyclohexane + cyclohexene (P93%), with small quantities of benzene (in
the range 17%) that increased at higher temperatures. It seems
that for these HY loading used in these catalysts there is not significant isomerisation activity to produce methylcyclopentane even
though they presented high HDO activity. Similarly, NiMo sulde
catalyst tested at 523 K showed a large selectivity toward cyclohexane (50%) and benzene (46%), and a low selectivity toward
methylcylopentane (3%). Comparison of the selectivity of the Pdbased catalysts with that of the sulde NiMo catalyst indicates a
large benzene formation on the latter catalyst which in turn is
slightly higher than that of Pd/HY counterpart (46% vs. 38%). Taking into account the high cost of hydrogen, a large benzene formation should be considered as a positive feature of the Pd/HY
catalyst. Contrary to our study, direct hydrogenolysis of phenol
to benzene was inhibited on acidic Pd/C catalysts having the combination of a hydrogenation catalyst and a strong acid, which were
tested in aqueous-phase HDO of bio-derived phenols to cycloalkanes [31,32]. However, the comparison of the activity results reported in literature for supported Pd catalysts is difcult, if not
impossible, because different reactors, reaction conditions, solvents and different catalyst formulations have been employed.
Main products obtained on the Pd/Al catalyst are cyclohexane
and cyclohexene, and upon increasing temperature higher
amounts of benzene were formed. This is likely due to the fact
that the benzenecyclohexenecyclohexane equilibrium suffers

signicant thermodynamic limitations depending on the temperature (at higher temperatures benzene formation is favoured). Another reason could be the variation of the hydrogenation activity
of the catalyst sites associated to their deactivationmodication
along the time-on-stream. Similarly to Pd/Al, Pd/HY catalyst shows
strong selectivity-temperature dependence. On the contrary, the
stability in the product distribution of the Pd/20%HYAl catalyst
is remarkable even if it suffered some deactivation along the tests.
Another interesting comparison is the product yields obtained
at 523 K for the Pd-based catalysts and the sulded NiMo/Al2O3
zeolite one (Fig. 8(b)). Both Pd/20%HYAl and sulded NiMo catalysts produced similar yields of O-free products. However, the Pd/
HY catalyst produced exclusively cyclohexene and cyclohexane
whereas the reference one yields equal amount of benzene and
cyclohexene + cyclohexane. Thus, it can be concluded that less
hydrogen is required for phenol transformation over commercial
NiMo catalyst than on the Pd/20%HYAl one. Finally, the Pd/HY
catalyst showed the lowest yield of all O-free products linked with
the lowest phenol conversion on this catalyst.
3.4. Catalyst deactivation by coke
In order to shed light on the origin of catalyst deactivation during phenol HDO over reduced Pd catalysts, the quantity of coke
deposited was evaluated by TGA analysis. The amount of coke
formed in the spent catalysts after 16 h on stream was evaluated
by temperature-programmed oxidation in a 20% O2/N2 mixture.
The weight change during oxidation together with DTG proles
are shown in Fig. 9. Regardless of the support, all DTG proles
show one intense peak at temperatures below 773 K, which is
due to burning of carbonaceous residues. For the Pd/10%HYAl
and Pd/20%HYAl catalysts, this peak shows a shoulder located
on the higher temperature region indicating the presence of more
polymerized-type of coke species. On the contrary, the DTG peak of
spent Pd/HY catalyst shows some shoulder at lower temperatures
suggesting the formation of carbonaceous residues with higher degree of hydrogenation, which are easier to burn off. The percentages of total mass loss corresponding to all types of burnt coke
species follows the trend: Pd/HY (16.3%)  Pd/20%HYAl (8.4
%) > Pd/10%HYAl (7.1%) > Pd/Al (6.4%).
3.5. Catalyst activity-structure correlation
In this study, mixed aluminazeolite HY was successfully used
as supports for the catalytic response of palladium catalysts in
the phenol HDO reaction. Under the reaction conditions employed

1070

S. Echeandia et al. / Fuel 117 (2014) 10611073

Pd/Al

100

Pd/10%HY-Al

100

0.00

Coke: 6.4%

Coke: 7.1%

95

90
-0.04

-0.02

90
-0.04

dw/dT [mg/C]

-0.02

Mass loss (%)

95

dw/dT [mg/C]

Mass loss (%)

0.00

85

85
-0.06

436 C

-0.06
438 C

80

200

400

600

800

80

1000

200

Pd/20%HY-Al

100

400

600

800

1000

Temperature (C)

Temperature (C)

Pd/HY

100

0.00

-0.02

90
-0.04

85

-0.05

95

Coke: 16.3%
-0.10
90
-0.15

dw/dT [mg/C]

Coke: 8.4%

Mass loss (%)

95

dw/dT [mg/C]

Mass loss (%)

0.00

85
-0.20

-0.06
502 C

80

200

400

600

800

1000

Temperature (C)

80

200

400

600

800

-0.25
1000

Temperatue (C)
Fig. 9. TPO/TGA proles of spent palladium catalyst.

(T = 523623 K, P = 15 bar, WHSV = 0.5 h1), the best catalytic response was achieved on the Pd catalyst supported on mixed
20%HY80%Al2O3 carrier. At a rst glance, three interesting observations can be put forward to describe the most interesting features of this work:
(1) Upon the reaction conditions studied, the phenol transformation over Pd/Al2O3, Pd/HYAl2O3 and Pd/HY catalysts
using n-octane as solvent proceeds according to the same
reaction routes as those observed for sulded NiMo/Al2O3
catalysts [1419]: (i) direct hydrogenolysis to benzene, and
(ii) hydrogenation to cylohexanol followed by direct cyclohexanol hydrogenolysis to cyclohexane.
(2) All catalysts studied shows 100% selectivity toward O-free
products
(3) The large extent of benzene formation on the Pd/HY catalysts suggests that the excessive H2-consumption could be
avoided using HY carrier. This catalyst shows hydrocracking
and isomerization functions, as revealed by the methylcyclopentane formation.
Considering the catalysts characterization data, several explanations can be advanced to explain the greater intrinsic activity
of the Pd/20%HYAl catalyst in the HDO of phenol at steady-state:
(i), the optimized support acidity, as evidenced by TPD-NH3

measurements, which led to an increase of the catalyst activity


without excessive coking; (ii), an increase of specic BET surface
area; (iii), additional activity from acid sites [60], and (iv) a large
palladium dispersion, as evidenced by HRTEM and TPR results.
The catalysts showing the best Pd dispersion, showed the largest
activity per palladium surface atom. This suggests that the phenol
HDO reaction could be a structure sensitive reaction, i.e. specic
activity might to depend on the details of the surface structure of
the palladium crystallites.
Contrary to the basic supports, H-spillover is a common feature
observed on Pd catalysts supported on acidic substrates which contributes to enhance HDO activity [25]. This is a key requirement
needed for the effective CO bond cleavage of O-containing compounds. In this sense, a direct correlation between the support
acidity and the HDO activity of Rh catalysts supported on different
carriers was observed [25]. Since Rh/ZrO2 sample was found to be
much less active than its counterparts supported on more acidic
substrates (alumina, silicaalumina and nitric-acid treated carbon
black), the authors concluded that bifunctional catalysts are
needed for the effective O-removal from phenol.
It is generally accepted that the HDO reaction over noble metals
supported on acidic substrates occurs on metal sites as well as on
acid sites being H2 dissociated on metal sites, while O-containing
compounds are adsorbed and activated on either metal sites or
on the cations/oxygen vacancies located at the metal-support

1071

S. Echeandia et al. / Fuel 117 (2014) 10611073


Table 3
Acid properties of pure supportsa and reduced Pd/Al, Pd/HYAl and Pd/HY catalysts as
determined by TPD-NH3.
Catalyst

Pd/Al
Pd/10%HYAl
Pd/20%HYAl
Pd/HY
a

Amount of acid sites mmol NH3 g1


cat
Weak T < 523 K

Moderate T = 523773 K

Total

0.9
0.9
1.8
1.8

0.4
0.8
0.7
3.0

1.3
1.7
2.5
4.8

(0.9)
(1.8)
(1.8)
(3.4)

(1.1)
(1.9)
(3.6)
(7.7)

(2.0)
(3.7)
(5.4)
(11.1)

Data of the bare supports are given in parenthesis.

interface [1419]. In this work, the possible contribution of the


H-spillover to the overall HDO activity should be obtained from a
correlation between the catalyst acidity and HDO activity. Indeed,
with exception of Pd/HY, the total acidity (Table 3) and HDO
activity (Table 5) of the catalysts follows the same trend: Pd/
20%HYAl > Pd/10%HYAl > Pd/Al. The activity and selectivity results reported in this work are in good agreement with literature
reports on the Pd catalysts supported on acidic carriers indicating
that a bifunctional mechanism is involved in HDO reaction over all
Pd-based catalysts. This is also conrmed by analyzing products of
HDO reaction over Pd catalysts supported on non-acidic carriers
[41,42]. For example, in the vapour phase hydrogenation of phenol
over Pd supported on non-acidic substrates (alumina and LTL zeolite), the major products detected in the reaction at T = 523 K and
20 bar H2 pressure were cyclohexanone and cyclohexanol with
cyclohexene, cyclohexane and benzene as minor products [42].
Thus, the catalysts supported on non-acidic carriers are good supports if the objective of HDO reaction is cyclohexanone. The ratedetermining step of the target reaction over Pd catalysts supported
on non-acidic carriers is probably a surface reaction between the
strongly bond phenol and the weakly adsorbed H-atom, as it was
derived from a kinetic study of phenol hydrogenation over Pd/
MgO catalysts [41].

In good agreement with recent reports on the supported Pd


HDO catalysts [1419 and references within], product distribution
recorded in the present work suggests that Pd metal sites can be
responsible of phenol hydrogenation to cyclohexanol followed by
its transformation into cyclohexane and nally dehydrogenated
to form benzene. The Pd/HY catalyst tested at 523 K shows exceptionally high selectivity to benzene (38%) indicating a large O-elimination from phenol via hydrogenolysis reaction route (see
Scheme 1). This result appears quite interesting because such reaction required less hydrogen consumption. Considering the highest
density of Pd sites on the surface of Pd/HY sample (Table 4) and its
largest acidity among the catalysts studied (Table 3), it appears
reasonably assume that the couple of Pd and acid sites in close
vicinity are responsible for the higher methylcylopentane formation observed using this catalyst. However, hydrocracking and
isomerization are not necessary catalyst functions because total
oxygen elimination was achieved for all catalysts studied.
Finally, TPO experiments conrmed that the main cause of catalyst deactivation in phenol HDO is deposition of carbonaceous
residues/coke on the catalyst surface showing the Pd/HY catalyst
the largest amount of coke among the catalysts studied. For this
catalyst, an increase of the selectivity toward benzene formation
with an increase of reaction temperature is accompanied by a
strong decrease of the phenol conversion (Table 5). Considering
the large amount of coke formed (Fig. 9), the increase of the selectivity toward benzene is due to a decrease of the catalyst hydrogenation ability linked with coke formation on the metal sites. The
amount of coke is plotted in Fig. 10 against the catalyst total acidity. From this gure it is obvious that the catalyst acidity is a decisive factor inuencing on coke formation, as expected [61,62].
Thus, the larger extent of coke formed on the Pd/HY catalyst could
be due to its larger surface concentration of acid sites. Thus,
although the Pd/HY catalyst can be effective for the HDO reaction,
it can be concluded that the density of its acid sites needs to be
optimized; large enough to achieve high conversion of phenolic

Table 4
Binding energy (eV) and surface atomic ratios (from XPS), and HRTEM parameters (statistics analysis) of the reduced (573 K) Pd/Al, Pd/HYAl and Pd/HY catalysts.
Catalyst

Pd/Al
Pd/10%HYAl
Pd/20%HYAl
Pd/HY

XPS

TEM

Pd 3d5/2 (eV)

Si/(Si + Al) at

Pd/(Si + Al) at

Particle size (nm)

Surface density of Pd0 particles per 50 nm2

334.5
334.9
334.9
335.0

0.044
0.092
0.654

0.0140
0.0139
0.0084
0.0024

5.7 2.2
5.0 1.7
4.6 1.2
3.2 0.7

26 5
11 2
20 6
48 6

Table 5
Inuence of temperature on the phenol conversion and selectivity in HDO of phenola over Pd/Al, Pd/HYAl, Pd/HY catalysts.
Catalyst

a
b

T (K)

Conv. (%)

Selectivity (%)
Benzene

Cyclohexene + cyclohexane

Methyl-cyclo-pentane

Pd/Al

523
573
623

28
35
22

0
2
13

100
98
85

0
0
2

Pd/10%HYAl

523
573
623

21
22
29

1
2
4

99
97
95

0
1
1

Pd/20%HYAl

523
573
623

63
57
46

4
7
7

95
93
93

1
0
0

Pd/HY

523
573
623

11
20
35

38
25
42

62
58
25

0
17
33

Sulded NiMob

523

62

46

50

Reaction conditions were: T = 523623 K, P = 1.5 MPa, WHSV = 0.5 h1.


A commercial NiMo/Al2O3zeolite sulde catalyst (HCK).

1072

S. Echeandia et al. / Fuel 117 (2014) 10611073

20

Coke (%)

15

0.9

R=

10

5
-1

Total catalyst acidity (mmol NH3 gcat )


Fig. 10. Inuence of total acidity of the oxide precursors (from TPD-NH3) on the
coke formation in HDO of phenol over supported Pd catalysts (T = 250-350 C,
P = 1.5 MPa and WHSV = 0.5 h1).

compounds, but not too high that would result in fast deactivation.
The lowering of catalyst activity due to undesirable oxygenation
function of residual chloride ions could be excluded since both
XPS and EDX/SEM techniques did not detect the chloride impurity
in the reduced catalysts.
4. Conclusions
 As a general conclusion, the use of mixed HYAl2O3 (20 wt.% of
HY) as support is benecial for HDO reaction over palladiumbased catalysts.
 Conversion results showed that, upon reaction conditions
employed (ow reactor, T = 523 K, PH2 = 15 bar, solvent:
n-octane; WHSV = 0.5 h1), the HDO of phenol over Pd/
20%HYAl catalyst led to similar yields of O-free products as
those obtained using the sulded NiMo/Al2O3zeolite hydrocracking catalyst.
 The largest activity of Pd/20%HYAl catalyst in the HDO of phenol
was explained as due to the hydrogen spillover phenomenon and
largest active phase dispersion on the carrier surface. Catalysts
further developed from this one could be used in the nal stage
of bio-liquid transformation into high quality and O-free bio-oil.
 A drastic decrease in activity going from Pd/20%HYAl to Pd/HY
sample indicated that the high catalyst acidity would result in
fast catalyst deactivation.

Acknowledgements
Financial support by the Community of Madrid (Spain) and
European Union (Project S2009/ENE-1743) is gratefully acknowledged. S.E. acknowledges nancial support from the Spanish Ministry of Science and Innovation (ENE2010-21198-C04-03) and the
Basque Autonomous Government (IE09-263-IE10-288-VALCAPEF).
The authors acknowledge the technical assistance of Dr. M.A. Pea
in TPO/TG measurements.
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.fuel.2013.10.011.
References
[1] Bridgwater AV, Meier D, Radlein D. An overview of fast pyrolysis of biomass.
Org Geochem 1999;30(12):147993.
[2] Demirbas A. Biomass resource facilities and biomass conversion processing for
fuels and chemicals. Energy Convers Manage 2001;42(11):135778.

[3] Sarkar S, Kumar A. Large-scale biohydrogen production from bio-oil. Bioresour


Technol 2010;101:735061.
[4] Mohan D, Pittman Jr CU, Steele PH. Pyrolysis of wood/biomass for bio-oil: a
critical review. Energy Fuels 2006;20:84889.
[5] Grange P, Laurent E, Maggi R, Centeno A, Delmon B. Hydrotreatment of
pyrolysis oils from biomass: reactivity of the various categories of oxygenated
compounds and preliminary techno-economical study. Catal Today
1996;29:297301.
[6] Oasmaa A, Kuoppala E, Solantausta Y. Fast pyrolysis of forestry residue. 2.
Physicochemical composition of product liquid. Energy Fuels 2003;17:43343.
[7] Furimsky E. Catalytic hydrodeoxygenation. Appl Catal A: Gen
2000;199:14790.
_ Viljava T-R, Krause AOI. Hydrodeoxygenation of methyl esters on
[8] Senol OI,
sulphided NiMo/gamma-Al2O3 and CoMo/gamma-Al2O3 catalysts. Catal Today
2005;106:1869.
[9] Laurent E, Delmon B. Study of the hydrodeoxygenation of carbonyl, carylic and
guaiacyl groups over sulded CoMo/c-Al2O3 and NiMo/c-Al2O3 catalyst: II.
Inuence of water, ammonia and hydrogen sulde. Appl Catal A
1994;109:97115.
[10] Ferrari M, Bosmans S, Maggi R, Delmon B, Grange P. CoMo/carbon
hydrodeoxygenation catalysts: inuence of the hydrogen sulde partial
pressure and of the suldation temperature. Catal Today 2001;65:25764.
_ Viljava T-R, Krause AOI. Effect of sulphiding agents on the
[11] Senol OI,
hydrodeoxygenation of aliphatic esters on sulphided catalysts. Appl Catal A:
Gen 2007;26:23644.
[12] Viljava T-R, Komulainen RS, Selvam T, Krause AOI. Stability of CoMo/Al2O3
catalysts: effect of HDO cycles on HDS. Stud Surf Sci Catal 1999;127:14552.
[13] Navarro RM, Pawelec B, Castao P, lvarez-Galvn MC, Loricera CV, Fierro JLG.
Upgrading of bio-liquid on different mesoporous silica-supported CoMo
catalysts. Appl Catal, B 2009;92:15467.
[14] Huber GW, Iborra S, Corma A. Synthesis of transportation fuels from biomass:
chemistry, catalysts, and engineering. Chem Rev 2006;106:404498.
[15] Elliot DC. Historical developments in hydroprocessing bio-oils. Energy Fuels
2007;21:1792815.
[16] Mortensen PM, Grunwldt J-D, Jensen PA, Knudsen KG, Jensen AD. A review of
catalytic upgrading of bio-oil to engine fuels. Appl Catal A 2011;407:119.
[17] He Z, Wang X. Hydrodeoxygenation of model compounds and catalytic
systems for pyrolysis bio-oils upgrading. Catal Sustain Energy 2012;1:2852.
[18] Butler E, Devlin G, Meier D, McDonnell K. A review of recent laboratory
research and commercial developments in fast pyrolysis and upgrading.
Renew Sustain Energy Rev 2011;15(8):417186.
[19] Bu Q, Lei H, Zacher AH, Wang L, Ren S, Liang J, et al. A review of catalytic
hydrodeoxygenation of lignin-derived phenols from biomass pyrolysis.
Bioresour Technol 2012;124:4707.
[20] Elliott DC, Hu J, Hart TR, Neuenschwander GG. US Pattent WO 2008/151269
A2.
[21] Lin Y-C, Li C-L, Wan H-P, Lee H-T, Liu C-F. Catalytic hydrodeoxygenation of
guaiacol on Rh-based and sulded CoMo and NiMo catalysts. Energy Fuels
2011;25:8905.
[22] Gutierrez A, Kaila RK, Honkela ML, Slioor R, Krause AOI. Hydrodeoxygenation
of guaiacol on noble metal catalysts. Catal Today 2009;147:23946.
[23] Ardiyanti AR, Gutierrez A, Honkela ML, Krause AOI, Heeres HJ. Hydrotreatment
of wood-based pyrolisis oil using zirconia-supported mono- and bimetallic (Pt,
Pd, Rh) catalysts. Appl Catal A 2011;407:5666.
[24] Veriansyah B, Han JY, Kim SK, Hong S-A, Kim YJ, Lim JS, et al. Production of
renewable diesel by hydroprocessing of soybean oil: effect of catalysts. Fuel
2012;94:57885.
[25] Lee CR, Yoon JS, Suh Y-W, Choi J-W, Ha J-M, Suh DJ, et al. Catalytic roles of
metals and supports on hydrodeoxygenation of lignin monomer guaiacol.
Catal Commun 2012;17:548.
[26] Pham T, Lobban LL, Resasco DE, Mallinson RG. Hydrogenation and
hydrodeoxygenation of 2-methyl-2-pentenal on supported metal catalysts. J
Catal 2009;266:914.
[27] Sitthisa S, Resasco DE. Hydrodeoxygenation of furfural over supported metal
catalysts: a comparative study of Cu, Pd and Ni. Catal Lett 2011;141:78491.
[28] Sitthisa S, Pham T, Prasomsri T, Sooknoi T, Mallinson RG, Resasco DE.
Conversion of furfural and 2-methylpentanal on Pd/SiO2 and PdCu/SiO2
catalysts. J Catal 2011;280:1727.
[29] Yu W, Tang Y, Mo L, Chen P, Lou H, Zheng X. Bifunctional Pd/Al-SBA-15
catalyzed one-step hydrogenationesterication of furfural and acetic acid: a
model reaction for catalytic upgrading of bio-oil. Catal Commun
2011;13(1):359.
[30] Kikhtyanin OV, Rubanov AE, Ayupov AB, Echevsky GV. Hydroconversion of
sunower oil on Pd/SAPO-31 catalyst. Fuel 2010;89:308592.
[31] Zhao C, Kou Y, Lemonidou AA, Li XB, Lercher JA. Highly selective catalytic
conversion of phenolic bio-oil to alkanes. Angew Chem Int Ed
2009;48:398790.
[32] Zhao C, He J, Lemonidou AA, Li X, Lercher JA. Aqueous-phase
hydrodeoxygenation of bio-derived phenols to cycloalkanes. J Catal
2011;280:816.
[33] Parsell TH, Owen BC, Klein I, Jarrell TM, Marcum CL, Haupert LJ, et al. Cleveage
and hydrogenation (HDO) of CO bonds relevant to lignin conversion using Pd/
Zn synergetic catalysts. Chem Sci 2013;4:80613.
[34] Wildschut J, Mahfud FH, Venderbosch RH, Heeres HJ. Hydrotreatment of fast
pyrolysis oil using heterogeneous noble metal catalysts. Ind Eng Chem Res
2009;48:1032434.

S. Echeandia et al. / Fuel 117 (2014) 10611073


[35] Sun J, Karim AM, Zhang H, Kovarik L, Shari Li X, Hensley AJ. Carbon-supported
bimetallic PdFe catalysts for vapor-phase hydrodeoxygenation of guaiacol. J
Catal 2013;306:4757.
[36] Snre M, Kubickov I, Mki-Arvela P, Chichova D, Ernen K, Yu D. Catalytic
deoxygenation of unsaturated renewable feedstocks for production of diesel
hydrocarbons. Fuel 2008;87:93345.
ejka J. Hydrodeoxygenation of
[37] Bejblov M, Zmostny P, Cerveny L, C
benzophenone on Pd catalysts. Appl Catal A 2005;296:16975.
ejka J.
[38] Prochzkov D, Zmostny P, Bejblov M, Cerveny L, C
Hydrodeoxygenation of aldehydes catalyzed by supported palladium
catalysts. Appl Catal A 2007;332:5664.
[39] Elliot DC, Hart TR. Catalytic hydroprocessing of chemical models for bio-oil.
Energy Fuels 2009;23:6317.
[40] Tang Z, Lu Q, Zhang Y, Zhu X, Guo Q. One step bio-oil upgrading through
hydrotreatment, esterication, and cracking. Ind Eng Chem Res
2009;48:696929.
[41] Mahata N, Vishwanathan V. Kinetics of phenol hydrogenation over supported
palladium catalyst. J Mol Catal A: Chem 1997;120:26770.
[42] Talukdar K, Bahattacharyya KG, Sivasanker S. Hydrogenation of phenol over
supported platinum and palladium catalysts. Appl Catal A 1993;96:22939.
[43] Nimmanwudipong T, Runnebaum RC, Block DE, Gates BC. Cyclohexanone
conversion catalyzed by Pt/gamma-alumina: evidence of oxygen removal and
coupling reactions. Catal Lett 2011;141:10728.
[44] Zhu X, Lobban LL, Mallinson RG, Resasco DE. Bifunctional transalkylation and
hydrodeoxygenation of anisole over a Pt/HBeta catalyst. J Catal
2011;281(1):219.
[45] Hong D-Y, Agrawal PK, Miller SJ, Jones CW. Hydrodeoxygenation of phenol
over zeolite-supported metal catalysts. 2009 AIChE annual meeting. Catalysis
and Reaction Engineering Division.
[46] Hong D-Y, Miller SJ, Agrawal PK, Jones CW. Hydrodeoxygenation and coupling
of aqueous phenolics over bifuncional zeolite-supported metal catalysts. Chem
Commun 2010;46:103840.
[47] Navarro RM, Pawelec B, Trejo JM, Mariscal R, Fierro JLG. Hydrogenation of
aromatics on sulfur-resistant PtPd bimetallic catalysts. J Catal
2000;189:18494.
[48] Egia B, Cambra JF, Gemez B, Arias PL, Pawelec B, Fierro JLG. Hydrocracking
activity of NiMo-USY zeolite hydrotreating catalysts. Stud Surf Sci Catal
1997;106:56771.

1073

[49] Gregg SJ, Sing KSW. Adsorption, surface area and porosity. 2nd ed. New
York: Academic Press; 1982.
[50] Leofanti G, Padovan M, Tozzola G, Venturelli B. Surface area and pore texture of
catalysts. Catal Today 1998;41:20719.
[51] Lin B, Wang R, Yu X, Lin J, Xie F, Wei K. Physicochemical characterization and
H2-TPD study of alumina supported ruthenium catalysts. Catal Lett
2008;124:17884.
[52] Okal J, Zawadzki M. Inuence of catalyst pretreatments on propane oxidation
over Ru/c-Al2O3. Catal Lett 2009;132:22534.
[53] Luo MF, Zheng XM. Redox behaviour and catalytic properties of Ce0.5Zr0.5O2supported palladium catalysts. Appl Catal A 1999;189:1521.
[54] Chang TC, Chen JJ, Yeh CT. Temperature-programmed reduction and
temperature-resolved sorption studies of strong metal-support interaction in
supported palladium catalysts. J Catal 1985;96:517.
[55] Vannice MA, Chou P. Benzene hydrogenation over supported and unsupported
palladium: I. Kinetic behavior. J Catal 1987;107(1):12939.
[56] Lieske H, Vlter J. Palladium redispersion by spreading of palladium(II) oxide
in oxygen treated palladium/alumina. J Phys Chem 1985;89(10):18412.
[57] Gopinath R, Babu NS, Kumar JV, Lingaiah N, Sai Prasad PS. Inuence of Pd
precursor and method of preparation on hydrodechlorination activity of
alumina supported palladium catalysts. Catal Lett 2008;120:3129.
[58] Pawelec B, La Parola V, Thomas S, Fierro JLG. Enhancement of naphthalene
hydrogenation over PtPd/SiO2Al2O3 catalyst modied by gold. J Mol Catal A:
Chem 2006;253:3043.
[59] Murcia-Mascars S, Pawelec B, Fierro JLG. Aromatics hydrogenation on PtPd
metals supported on Zr-phosphate. Catal Commun 2002;3:30511.
[60] Pawelec B, Mariscal R, Navarro RM, van Bokhorst S, Rojas S, Fierro JLG.
Hydrogenation of aromatics over supported PtPd catalysts. Appl Catal A
2002;225:22337.
[61] Arias PL, Zugazaga F, Gemez B, Cambra JL, Pawelec B, Fierro JLG.
Hydrodesulfurization of diesel oil on Pt and NiW phases supported on
aluminazeolite mixtures. In: Symposium on recient advances in heteroatom
removal presented before the Division of Petroleum Chemistry, Inc. 215th
National meeting, Americam Chemical Society, Dallas, TX, March 29April 3,
1998. ACS symposium series, vol. 43; 1998. p. 637.
[62] Castao P, Elordi G, Olazar M, Aguayo AT, Pawelec B, Bilbao J. Insights into the
coke deposited on HZSM-5, Hb and HY zeolites during the cracking of
polyethylene. Appl Catal B 2011;104:91100.

You might also like