Sie sind auf Seite 1von 12

Applied Catalysis A: General 510 (2016) 98109

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Improving the selectivity in hydrocracking of phenanthrene over


mesoporous Al-SBA-15 based FeW catalysts by enhancing
mesoporosity and acidity
Jonatan R. Restrepo-Garcia, Vctor G. Baldovino-Medrano, Sonia A. Giraldo
Centro de Investigaciones en Catlisis (@CICATUIS), Escuela de Ingeniera Qumica Cra 27, Calle 9, Universidad Industrial de Santander (UIS) Bucaramanga,
Colombia

a r t i c l e

i n f o

Article history:
Received 25 June 2015
Received in revised form 21 October 2015
Accepted 31 October 2015
Available online 15 November 2015
Keywords:
Hydrocracking
Molecular sieves
FeW catalysts
Phenanthrene
Al-SBA-15

a b s t r a c t
Catalysts for heavy oil hydrocracking require an enhanced mesoporosity (higher pore diameters) and a
moderate acidic function (mild acidity) to treat the bulky molecules present in this kind of feedstock and
to yield middle distillates (MD). In this work, we have synthesized ve different kinds of mesoporous
silica based on SBA-15 by modifying some of the variables of their synthesis with the aim at enhancing
mesoporosity. NH4 F and 1,3,5-trimethylbenzene (TMB) were used to modify the mesostructured arrangement in SBA-15. TMB modied SBA-15 materials exhibited the highest textural properties (2.16 cm3 g1
and 17 nm) in comparison to NH4 F modied silica (1.90 cm3 g1 and 10.8 nm) and pristine SBA-15 silica
(1.03 cm3 g1 and 12 nm). The acidity of the SBA-15 based materials was modied by a post-synthesis
grafting procedure to Si/Al molar ratios of 10, 25, and 40. SBA-15 based materials modied with both
Al and TMB were used as supports for FeW suldes. These catalysts were tested in the hydrocracking
of phenanthrene. In general, all of the catalysts supported on Al-SBA-15 based materials were selective to the ring opening reaction of phenanthrene in contrast to the results obtained over a commercial
NiMo/-Al2 O3 catalyst. Such trend was associated to the presence of Brnsted acid sites on the surface
of the Al modied supports as shown 27 Al MAS NMR analysis.
2015 Elsevier B.V. All rights reserved.

1. Introduction
The increase in the demand of middle distillates (MD) production by rening heavy oils remains a challenging task [14]. The
efciency of conventional catalysts processing heavy oils is poor
due to diffusional limitations. Therefore, researchers have focused
on developing materials with larger pores to use them as supports
of the catalysts aiming the conversion of heavy oils into MD [57].
It is possible to synthesize and modify silica based materials
such as MCM-41 (Mobil Composition of Matter No 41), SBA-15
(Santa Barbara Amorphous No 5), and MCF (Mesocellular Foam),
and use them as catalytic supports for upgrading heavy oils [811].
In particular, SBA-15 is a type of mesoporous silica, which possesses wide pore volume, narrow pore size distributions, and better
hydrothermal stability as compared to, for example MCM-41 [12].
Therefore, SBA-15 is an interesting candidate for supported catalysts for hydroconversion of heavy oils [1214]. SBA-15 is formed

Corresponding author. Fax: +57 7 6344684.


E-mail address: sgiraldo@uis.edu.co (S.A. Giraldo).
http://dx.doi.org/10.1016/j.apcata.2015.10.051
0926-860X/ 2015 Elsevier B.V. All rights reserved.

by a hexagonal array of tubular and uniform cylindrical channels,


which is formed in strong acid media during the synthesis of the
material thanks to the micelles built over triblock copolymers [9].
Such copolymers act as both surfactant and a template during the
synthesis of the material [10,15,16]. Aiming pore size widening,
Zhao et al., studied the use of micellar swelling agents to modify
the textural and structural properties of SBA-15, obtaining silica
materials with larger pore sizes by using 1,3,5-trimethylbenzene
(TMB) [14,15]. It has also been reported that the use of organic cosolvents such as hexane or n-butanol can act as a swelling agents
during the synthesis of SBA-15, hence widening pore size [8,9,16].
On the other hand, an improvement on the thickness of SBA-15 was
reported after the addition of small quantities of inorganic salts during the synthesis of the material [1618], this feature makes SBA-15
a specic mesostructured material.
Another issue when trying to apply SBA-15 as catalytic support
for heavy oil hydroconversion is its modest acidity [19]. Such low
acidity is explained by the fact that the structure of SBA-15 consists
mainly of neutral Si atoms. To improve the acidity of SBA-15 based
supports, different metals such as Al, Ti, Zr, and Sn have been used
as promoters for the formation of Brnsted acid sites [2022]. In the

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

case of aluminum, aluminum loaded SBA-15 materials have been


prepared by varying the Si/Al molar ratio of the materials, aiming
to enhance the acidity of the materials [2326].
In this contribution, we investigated how to enhance both
the porosity and acidity of SBA-15 type materials in order to
prepare catalysts for the hydroprocessing of heavy crude oils.
On the other hand, the inuence of adding NH4 F and TMB
(1,3,5-trimethylbenzene) during the synthesis of SBA-15 on their
structure and porosity was analyzed. Five different kinds of SBA-15
based mesoporous silica were thus obtained. Conversely, the acidity of selected SBA-15 based materials was modied by aluminum
grafting. SBA-15 based catalytic supports of different Si/Al molar
ratios: 10, 25, and 40 were thus synthesized. The active phase of
the SBA-15 type catalysts was based on a combination of FeW. In
general, iron based active phases have received less attention from
researchers, despite of being cheaper than conventional Ni and Co
based catalysts [7,2731]. The synthesized catalytic materials were
tested in hydroprocessing of phenanthrene, as a model reaction for
processing vacuum gas oil (VGO). Catalysts were characterized by
different techniques: N2 physisorption to calculate textural properties, XRD to study the changes in the structure of the materials,
SEM to observe changes in morphology, 27 Al MAS NMR to identify
the Al species on the surface of the selected materials, and PAD to
study the nature of OH groups on the surface of the materials and
their overall acidity. Therefore, the effect of both, textural and acid
properties of the Al-SBA-15 based materials in hydroprocessing
reactions is given.

2. Experimental
2.1. Mesoporous silica synthesis
Five kinds of SBA-15 based materials were prepared by adapting the procedures employed in some literature reports [8,15,17].
Pluronic P123 (EO20 PO70 EO20 , Mav = 5800, Aldrich) was used as a
structural direction agent (SDA), and tetraethylortosilicate (TEOS,
98% Aldrich) as a silica source. Synthesis was conducted in strong
acid media conditions (pH < 1) in accordance to the following nominal molar ratio: 1.0 TEOS:0.0169 P123:4.42 HCl:186 H2 O. The effect
of NH4 F and TMB (1,3,5-trimethylbenzene) addition to the structure of SBA-15 was studied. NH4 F (99% Aldrich) and TMB (98%, Alfa
Aesar) were always added to the synthesis mixture before TEOS
incorporation.
In a typical synthesis of SBA-15, 10 g of P123 were completely
dissolved in an aqueous solution of 1.3 M HCl (37 vol.%, Merck) and
under stirring (500 rpm). Then, 24 mL of TEOS was added dropwise and kept under stirring for 1 h. Subsequently temperature was
increased to 311 K and thus kept during 24 h. The milky suspension
thus obtained was then transferred into a 500 mL PTFE vessel, and
placed inside a mufe furnace at 403 K for 24 h for the hydrothermal
treatment. Afterwards, the vessel was allowed to cool down. The
recovered solid product was ltered and washed with deionized
water, and left to dry at ambient conditions. The recovered powder was calcined, with a heating rate of 2 K/min in a ow of dry air
(100 mL/min) at 773 K during 6 h in order to remove the template.
NH4 F modied silica materials were synthesized by adding
0.10 g and 0.15 g of NH4 F to the synthesis mixture before TEOS
incorporation, those samples were labeled: SBA-F(0.10) and SBAF(0.15), the value in brackets indicates the amount of NH4 F added.
TMB modied silica materials were obtained adding the
required amount of TMB to get TMB:P123 mass ratios of 0.15 and
0.75 before dropwise TEOS incorporation and kept under stirring
for 1 h. Those materials were labeled: SBA-TMB(0.15) and SBATMB(0.75), the value in brackets indicates the TMB:P123 mass ratio.
The other steps remained the same as SBA-15 preparation.

99

2.2. Aluminum incorporation into the SBA-15 type silica


Surface acidity of the supports was obtained by an Al postsynthesis grafting procedure [29]. A 0.02 M ethanol (Merck, 98%)
solution of aluminum isopropoxide (C9 H21 AlO3, Merck 98%) was
used as aluminum source. Powdered samples (5 g) of SBA-15 and
SBA-TMB(0.75) were grafted by soaking them in the C9 H21 AlO3
solution, thence keeping the resulting suspension under constant
stirring (500 rpm) for 18 h. Grafting was performed at pH between
9.0 and 9.5 using NH4 OH (28 vol.%, Aldrich, 98%). Appropriate
amounts of the aluminum precursor were employed to obtain
materials with Si/Al molar ratios of 10, 25, and 40. The grafted
materials were ltered and washed with ethanol to remove residua
from the aluminum precursor. Then, the recovered materials were
calcined in dry air (100 mL/min) at 773 K for 6 h. Al-modied materials were labeled following the nomenclature: Al-SBA-15-x and
Al-SBA-TMB(0.75)-x, where x indicates the Si/Al molar ratio.

2.3. Catalysts preparation


FeW catalysts supported on -Al2 O3 (Protocatalyse), Al-SBA15-x and Al-SBA-TMB(0.75)-x were prepared by the successive
incipient wetness impregnation method. Fe(NO3 )3 9H2 O (98%,
Aldrich) and (NH4 )6 H2 W12 O40. H2 O (98%, Aldrich) aqueous solutions (30% in excess of the required amount to ll pore volume of
catalysts) were used as Fe and W precursors, respectively. After
each impregnation step, materials were dried at 393 K (heating
rate of 2 K/min) during 12 h, and then calcined in dry air ow
(100 mL/min) at 773 K during 6 h. Fe2 O3 and WO3 loadings were
3 wt.% and 15 wt.%, respectively. The thus obtained materials were
labeled as FeW/Al-SBA-15-x and FeW/Al-SBA-TMB(0.75)-x. All
catalysts were sieved down to a particle size range between 25 m
and 75 m.
2.4. Characterization of supports and catalysts
2.4.1. Textural characteristics
The BET specic surface area (SBET , m2 g1 ) of the materials was
determined by N2 adsorptiondesorption isotherms measured at
77 K in a 3FLEX apparatus from Micromeritics.
Samples of 0.2 g of the materials were degassed in vacuum (3 Pa)
at 393 K for 12 h before the analysis.
SBET was calculated from data taken in the range of relative
pressures (P/P0 ) between 0.04 and 0.24 according to IUPAC recommendation [32]. Pore size distributions were estimated from
N2 -DFT models provided by Micromeritics depending on the geometry of the pores. For SBA-15 based materials, a model based on
cylindrical pores oxide surface, and for SBA-TMB(0.75) catalysts
a model based on slit pores was employed. Total pore volumes
were calculated from the amount of nitrogen adsorbed at a relative
pressure (P/P0 ) of 0.9.
2.4.2. XRD patterns
XRD patterns of the materials were recorded in a Bruker Advance
diffractometer with Da Vinci geometry using Ni-ltered CuK1
radiation (40 kV, 30 mA) instrument. Samples (0.2 g) were grinded
in an agate mortar down to a particle size of 38 m. Low angle
spectra were recorded to identify changes in the material ordering pattern. 2 range was scanned between 0.5 and 8 with a
step size of 0.01526 , and a counting time of 0.4 s per step. Qualitative analysis of the observed peaks was carried out by comparing
the observed prole with the diffraction prole reported in the
database PDF-4 + ICDD (SiO2 -00-058-0344). Wide angle analyses
were performed for all catalysts in order to conrm the occurrence

100

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

Fig. 1. N2 adsorptiondesorption isotherms, (a) silica materials, (b) catalysts.

of the respective oxidized species of the metals incorporated. In


this case, 2 range was scanned between 10 and 80 .

2.4.3. Scanning-electron-microscopy (SEM)


SEM measurements were made in a FEI Quanta 650 FEG apparatus operated at an electron voltage of 4.00 kV. Magnications of
3 m, and 500 nm were made, as depending on the size of the
particle to be observed.

2.4.4. 27 Al MAS NMR measurements


Solid state 27 Al MAS NMR measurements were conducted on
a Bruker ADVANCE 400 WB III instrument at B0 = 9.4 T. The resonance frequency corresponding to 27 Al was 104.3 MHz. All grafted
supports labeled as Al-SBA-15-x and Al-SBA-TMB(0.75)-x were
analyzed. The MestReNova 9.0 software was used to compute the
content of the incorporated aluminum species by a Gaussian integration of their respective peaks.

2.4.5. Proton afnity distributions (PAD)


Proton afnity distributions were computed from potentiometric titration performed in a TitroLine 7000 (SI Analytics) titrator.
Samples were grinded and sieve ddown to a particle size below
75 m [28]. Samples of 0.05 g of the materials were added to a
0.01N aqueous solution of NaNO3 (99.5% Merck). The suspension
was stirred for 30 min before performing titration. A 0.1 M aqueous
solution of NaOH (Merck 99%) was used for titrations in the basic
pH range. A volume of 0.03 mL of NaOH was added every 90 s until
reaching a pH of 10. Titration in the acidic pH range for a fresh sample was performed with a 0.1 M HCl (37 vol.%, Merck) until reaching
a pH of 3. The acquired data were used to construct a proton consumption function (f(log k)) as a function of pH by proton balance
[3335].
Alumina acid sites distribution was proposed by Knzinger and
Ratnasamy, representing the OH groups conguration on its surface
[35]. Hence, proton afnity distributions were analyzed taking into
account alumina acid sites contribution to Al modied silica materials. The quantication of the identied OH groups was performed
by deconvolution of the peaks from the proton consumption curves.
To this end, the OriginPro 8.5.1 software was used, indicating the
amount of acid sites [mmol H+ g1 ] present in the samples. Trends
in the quantication of the OH groups derived from the PADs show
mostly a qualitative value, being valid only for comparing the series
of catalysts prepared in this study.

2.5. Catalytic evaluation


Sulded catalysts were tested in a 570 mL stainless steel batch
reactor (Parr). The reactor was loaded with 250 mL of the reaction
feed composed of 4 wt.% phenanthrene (Aldrich, 98%), 94 wt.% nheptane (Merck, 98%); used as solvent and 2 wt.% n-hexadecane
(Aldrich, 98%) as an internal standard for chromatographic analyses. Before the reaction, catalysts (1 g) were activated ex-situ in a
ow of 100 mL min1 of a mixture of H2 S/H2 (15/85 vol.%) at 673 K
for 4 h.
The activated catalysts were added to the liquid feed of the
reactor, which was then hermetically sealed. Afterwards, H2 was
allowed into the reactor as to reach a PH2 = 2.2 MPa at room
temperature. Then, temperature was increased (heating rate of
2 K/min) up to the reaction temperature 623 K, and under agitation (1000 rpm). Once reaction temperature was reached, the
registered reactor pressure was 11 MPa, and the rst reaction
sample was collected. This sample was considered as corresponding to (zero) reaction time. Subsequently, samples were taken
each 15 min for the rst hour, each 30 min for the second hour,
and each hour for the last two hours. The volume of both reaction samples and its corresponding purges were measured every
time. Liquid reaction products of the reaction were analyzed by
GCMS, using a GC-HP 6890 instrument, equipped with an FID
detector and an HP-1 (100 m 0.25 mm 0.5 m), and an HP-5
(30 m 0.25 mm 0.25 m) columns, MS spectra were analyzed
by a comparison with the W.8.1 data basis, provided with the Data
Analysis Chemstation software from Agilent Technologies.
Commercial Catalyst NiMo/-Al2 O3 Protocatalyse was also
tested for comparison purposes.

2.6. Expression of catalytic results


2.6.1. Catalysts activity
Reaction rates were calculated by a rst order kinetic due to
the high excess of hydrogen [2,6,10,12,16,24]. Initial reaction rates
were thus computed by initial conversion data of phenanthrene
corresponding to t = 0 min until t = 60 min. The rst order Eq. (1)
proposed by Gevert and Otterstedt was used [36]. This equation
includes a correction factor due to the depletion in solution volume by repetitive purges and sampling during the experiments, as
depicted in Eq. (2).

ln

C 
i

C0

= kPhen Wf

t
V

(1)

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

t
V

n

ti ti1
i=1

101

(2)

Vi1

where Ci , C0 [g/mL] correspond to concentration of phenanthrene


at each t time when i samples are taken and at t = 0 min of reaction,
respectively, W [g] is the weight of the loaded catalyst, ti , ti1 [min]
is the time after taking two consecutive samples, and Vi1 [mL] is
the remaining liquid volume in the reactor after taking the i 1
sample.
2.6.2. Selectivity
Selectivity towards
tetrahidrophenanthrene,
Eq. (3).
yi =

9,10-dihydrophenanthrene, 1,2,3,4trans-stilben was computed from

moli
molPhenreaction

(3)

where moli are moles of i product (phenanthrene, trans-stilben,


9,10-dihydrophenanthrene, 1,2,3,4-tetrahydrophenanthrene) and
molPhenreaction : are moles of phenanthrene consumed during the
reaction.
3. Results
3.1. Characterization of catalysts and supports
3.1.1. Textural characteristics
Fig. 1 shows the N2 adsorptiondesorption isotherms for
all synthesized materials. Particularly, Fig. 1(a) shows results
for the synthesized silica materials, while Fig. 1(b) displays N2
adsorptiondesorption isotherms for the supported FeW oxide
precursors of the catalysts. In general, and according to the IUPAC
classication, all materials exhibited a type IV isotherm. This kind
of isotherm is typical of mesoporous materials [32].
Fig. 1(a) shows a wider H1 type hysteresis loop for the samples
SBA-15, SBA-TMB(0.15) and SBA-TMB(0.75). For this kind of hysteresis, the adsorption and desorption curves tend to be parallel
between them at high relative pressures [37]. Accordingly, these
materials possess tubular and uniform pores, as it is the case for
SBA-15 [15].
The isotherm for SBA-F(0.15) sample displayed a narrower and
not parallel H1 type hysteresis loop at high relative pressures.
Hence, it can be said that the addition of NH4 F during the synthesis
causes a loss in pore uniformity. This is conrmed by the absence
of the plateau at high relative pressures by increasing the amount
of NH4 F during the synthesis. For the case of SBA-F(0.10) silica, it is
evidenced that at high relative pressures, there was a partial deformation of the plateau as compared to SBA-15 silica, suggesting the
change in pore geometry at the same rate NH4 F amounts increased.
Fig. 1(b) shows narrower hysteresis loops for the supported
FeW materials as compared to the SBA-15 and SBA-TMB(0.75)
catalytic supports, when metal phases and aluminum species are
added. Therefore, porosity of the oxide precursor of the catalysts is
different from the one of the catalytic supports.
Fig. 2 shows pore size distributions for all the synthesized
mesoporous silica materials. Gaussian and narrow pore sizes distributions were evidenced for SBA-15 and silica materials modied
with TMB rather than the modied with NH4 F, indicating uniformity in pore sizes for TMB modied silica materials. In addition, all
materials exhibited the presence of pores in the range of microporosity.
Table 1 summarizes the calculated values of SBET (m2 g1 ), Smicro
2
(m g1 ), NSBET , the CBET constant, pore diameter (PD ), and pore
volume (PV ) for the synthesized materials. CBET values were positive

Fig. 2. Pore size distributions of SBA-15 based materials, (a) SBA-15, (b) SBATMB(0.75), (c) SBA-TMB(0.15), (d) SBA-F(0.10), (f) SBA-F(0.15).

and lower than 300, which evidence for an adequate tting of the
BET method for describing the recorded isotherms.
Considering the SBA-15 material, all synthesized materials
exhibited lower SBET values. Among the prepared materials, those
prepared with the addition of NH4 F had the lowest SBET . Such trend

102

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

Table 1
Physical properties determined from N2 sorption and XRD analysis.
Sample

SBET [m2 /g]

PDads [nm]

PV [cm3 /g]

CBET

NSBET

Smicro [m2 /g]

d(100) [nm]

WO3 crystal size (XRD) [nm]

-Al2 O3
SBA-15
SBA-F(0.10)
SBA-F(0.15)
SBA-TMB(0.15)
SBA-TMB(0.75)
FeW/Al-SBA-15-10
FeW/Al-SBA-15-25
FeW/Al-SBA-15-40
FeW/Al-SBA-TMB(0.75)-10
FeW/Al-SBA-TMB(0.75)-25
FeW/Al-SBA-TMB(0.75)-40

203
627.3
504.6
342.4
436.3
697.3
378.7
383.3
428.3
351.5
392.8
393.5

12
12
9.3
10.8
15.8
17.2
10.2
10.1
10.3
17.2
17.2
17.2

0.62
1.03
1.58
1.90
1.15
2.16
0.69
0.69
0.70
1.37
1.56
1.47

83
131
166
112
149
166
130
143
249
135
132
272

0.73
0.75
0.83
0.62
0.68
0.69

67.2
83.4
36.2
66.7
101.2
47.8
61.9
53.6
27.6
29.9
47.7

9.5
9.8
10.6
12.4
14.2
9.5
9.5
9.5
14.2
14.2
14.2

7.8
17.7

SBET : BET specic surface area, PV : pore volume, PDads : mesopore diameter corresponding to the maximum of the pore size distribution, CBET : constant BET, NSBET , (normalized
BET surface area) as dened in Eq. (4), Smicro : microporous area by t-plot, d(1 0 0): d spacing, WO3 crystal size as calculated by Scherrers equation.

allows to the collapse of mesostructured SBA-15 ordering pattern,


which is notorious due to the change of the shape in the hysteresis
loop, when adding NH4 F.
Pore diameter and pore volume increased for SBA-TMB(0.15)
and SBA-TMB(0.75) as the same rate TMB:P123 mass ratio
increased. Hence, the pore swelling effect of TMB was demonstrated. SBET area decreased as expenses at the increase of pore
diameter. For the samples, SBA-F(0.10) and SBA-F(0.15) materials, this increase in pore diameter became remarkable with the
highest amount of NH4 F. However, by comparing SBA-F(0.15) with
SBA-TMB(0.75), the effect of pore swelling on SBET was less in TMB
modied silica. Pore diameter and pore volume were the highest
for SBA-TMB(0.15). Consequently, this silica was selected for the Al
incorporation procedure with SBA-15 for comparison purposes.
For the oxide precursor of the catalysts (Table 1), there was
a decrease in SBET with decreasing the Si/Al molar ratio of the
materials. This decreasing is correlated with narrower hysteresis
loops in comparison to pristine SBA-15 and SBA-TMB(0.75) supports. Pore diameter reported values remained slightly constant for
FeW/Al-SBA-15-x catalysts than FeW/Al-SBA-TMB(0.75)-x catalysts. Therefore, as a result of metals impregnation and aluminum
content, a decrease in porosity for Al-modied supported catalysts
is evidenced by the decrease in the specic surface area at the same
rate Al content increases.
In terms of the changes of pore volume in the catalysts, a
decrease in the pore volume with a decrease in the Si/Al molar
ratio is evidenced as well. It may be directly related to the high-

est amount of Al species for the Si/Al molar ratio of 10, and it is
more evident for Al-SBA-15-x supported catalysts instead of AlSBA-TMB(0.75)-x supported catalysts, because of the higher pore
volume Al-SBA-TMB(0.75) supports possess.
To conrm the presence of nanoparticles of Fe and W oxides
inside the pores of the SBA-15 based catalysts, normalized SBET
values were calculated using the following equation [8]:

NSBET =

SBET catalysts
(1 x)
SBET supports

(4)

The normalized SBET is denoted as NSBET , and x is the weight


fraction of the metallic phases. If the metallic oxide component is
distributed at the support surface in the form of a close-packed
monolayer of a given thickness, no pore-blocking occurs. Therefore, the NSBET will decrease only as a result of the narrowing of
the mesopores. The introduction of Al species and Fe and W particles caused in a reduction of pore volume, which resulted in the
reduction of the normalized surface area. NSBET values increased
with increasing pore volumes.
The presence of microporosity is evidenced as well and reported
in Table 1 by t-plot (Harkins and Jura method) measured areas for
all samples.
All of the synthesized SBA-15 based materials showed higher
specic surface areas and porosity as compared to the commercial
-Al2 O3 support.

Fig. 3. Low angle XRD patterns, (a) silica materials, (b) oxide catalysts.

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

103

Fig. 4. Wide angle XRD patterns for FeW/Al-SBA-15-x and FeW/Al-SBA-TMB(0.75)-x oxide catalysts, x indicates Si/Al molar ratio.

3.1.2. XRD patterns


Fig. 3(a) and (b) depicts low angle diffraction proles for the
synthesized materials. For the SBA-15, Fig. 3(a), three diffraction
peaks, typical of this kind of materials, at 2 = 0.9, 1.5 and 1.9 corresponding to the (1 0 0), (1 1 0), and (2 0 0) were identied. Such
peaks are characteristic of the hexagonal symmetry of SBA-15 with
a space group p6mm [15].
For materials synthesized with NH4 F: SBA-F(0.10) and SBAF(0.15), XRD patterns (Fig. 3(a)) show that only the peak attributed
to the (1 0 0) plane is preserved. This peak is comparatively less
intense, wider, and is shifted to slightly lower 2 values. For TMB
modied materials, a 0.3 leftwards displacement of the 3 peaks is
observed for SBA-TMB(0.75), which shows a structural change and
rearrangement of the pores for the modied silica supports.
According to Table 1, the d spacing corresponding to the plane
(1 0 0) increases as the amounts of NH4 F and TMB increase, thus
indicating that the wall thickness of the silica materials decreased
as result of the increase in pore diameter.
Low angle diffraction patterns of Fe-W/Al-SBA-15 catalysts
(Fig. 3(b)) exhibited no changes as compared to the one obtained
over the SBA-15 material (Fig. 3(a)); their intensity and shape of the
diffraction peaks were very similar. This evidences that the addition
of aluminum by grafting as well as the impregnation of the FeW
metal did not destabilize the crystalline structure of the SBA-15
base material.
For catalysts prepared on Al-SBA-TMB(0.75)-x is possible to
observe in Fig. 3(b) a decrease in intensity of the three diffraction
peaks. It is more apparent for FeW/Al-SBA-TMB(0.75)-10 catalyst,
which possesses a greater amount of aluminum content. FeW/AlSBA-TMB(0.75)-40 catalyst showed almost the same diffraction
pattern than SBA-TMB(0.75) support (Fig. 3(a)). Based on the fact
that the metal content is the same for all six catalysts, the change
in the intensity is attributed to the aluminum incorporation for the
lower Si/Al molar ratio. The formation of Al species at higher rates
on the surface of the catalysts by decreasing the Si/Al molar ratio
diminished the intensity of the XRD signals. However, the ordering
pattern was also kept as the same Al-SBA-15-x catalysts, since the
original peaks position remained unaltered.
Fig. 4 shows the wide angle diffraction patterns for FeW/AlSBA-15-x and FeW/Al-SBA-TMB(0.75)-x oxide precursors of the
catalysts. FeW/Al-SBA-15-x catalysts showed only the characteristic hump of SiO2 material in the 2 range centered between 19
and 27 [8]. The high surface area of silica support favors dispersion

of the FeW phases, suggesting that either metals were homogeneously dispersed on the support, or their compositions were
below the detection limit of the X-ray signals (<4 nm). The absence
of XRD signals in wide angle region indicates that the particle size
of metals is below the coherence length of X-ray scattering.
Nevertheless, FeW/Al-SBA-TMB(0.75)-25 and FeW/Al-SBATMB(0.75)-40 catalysts exhibited other two peaks in 2 of: 33.2
and 24.4, which are attributed to poorly crystalline oxidized species
of Fe2 O3 and WO3 , respectively. The WO3 crystallite size as calculated by Scherrers equation is reported in Table 1, showing that at
higher Si/Al molar ratios, bigger WO3 particles are formed on the
surface of the Al-SBA-TMB(0.75) supports.
3.1.3. Scanning-electron-microscopy (SEM)
Results in Fig. 5 show the brous hexagonal morphology of the
SBA-15 material synthesized herein. Results are similar to those
previously reported in literature [15].
The effect on morphology by the addition of NH4 F in the synthesis of SBA-15 is illustrated in Fig. 6. The increase of NH4 F content

Fig. 5. SEM micrographs for SBA-15.

104

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

Fig. 6. SEM micrographs for SBA-F(0.15).

Fig. 8. PAD smoothed curves, (a) FeW/Al-SBA-15-x oxide catalysts, (b) FeW/AlSBA-15-TMB(0.75)-x oxide catalysts, x indicates Si/Al molar ratio.

to expand the pores [15,16]. Spherical particles observed resemble


MCF particles [3840]. Based on results obtained herein, it can be
stated that brous SBA-15 morphology was changed to spherical
morphology by the addition of TMB and NH4 F.

Fig. 7. SEM micrographs for SBA-TMB(0.75).

led to the formation of semi-spherical (peanut-shaped) particles.


Spherical particles are typical of amorphous silica, and here they
are associated to the collapse of the SBA-15 structure caused by the
pores swelling during synthesis. The interactions uoride ions with
the P123 micelles changes the latter copolymer in the formation of
well-developed silica particles as it has been previously stated by
several authors [8,9].
Fig. 7, featuring SEM pictures for SBA-TMB(0.75), shows the formation of well-dened spherical particles with an increase in the
TMB:P123 mass ratio. This result evidences that adding more quantity of TMB leads to swell pore diameter in SBA-15 as it has been
previously reported [11,15].
Recently, the 2D to 3D transition from SBA-15 to MCF (Mesocellular Foam) particles with the increase in the TMB in the mixture
synthesis was given in open literature [38]. In this case, spherical particles with different particle size were achieved, indicating
that there is an optimal mass ratio and a maximum load of TMB

3.1.4. Proton afnity distributions (PAD)


Fig. 8 shows PAD curves for the catalysts prepared with the AlSBA-15-x and Al-SBA-TMB(0.75)-x materials. Peaks associated to
type III (pH < 3.5) and IIA (3.5 < pH < 5.5) alumina acid sites, silanol
(pH 6.8), and a mixed group (silanol and siloxane, pH > 9) bonds for
both kinds of catalysts were evidenced.
PAD analysis of the SBA-15 based silica materials modied with
Al was performed considering the contributions of the alumina and
silica acid sites to the overall acidity and acid strength. Type III alumina acid sites refers to octahedral aluminum linked to hydroxyl
group, meanwhile Type IIA alumina acid sites refers to tetrahedral aluminum linked to the hydroxyl group on its surface [35].
The latter OH groups represent the strongest acid sites for alumina
supports.
PAD deconvolution as reported in Table 2, shows that the acidity
of the materials increases with the Si/Al molar ratio of the materials.
Similar results have been reported in literature for zeolites [41].
For the TMB supports there is an increase in the concentration of
silanol and of the mixed group. This group is similar to the IA group
in terms its acidity [42].
Al-SBA-15-40 and Al-SBA-TMB(0.75)-25 supports also presented a higher concentration of type III and type IIA acid sites in
comparison to other SBA-15 based supports grafted with Al. A leftwards displacement of 1.3 in pH of the peak associated to silanol
group (pH 6.8) for these supports was observed. Taking into account
that Brnsted acid sites are formed by the introduction of aluminum
in the terminal silanol bonds of the surface of silica, such a displacement evidences generation of Brnsted acid sites with increasing
the Si/Al molar ratio.
Regarding acid sites distribution of the oxide precursors of the
catalysts, an increase in the concentration of weak acid sites (pH > 9)

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

105

Table 2
PAD quantication of acid sites for oxide catalysts and silica materials.
Acid sites [mmol H+ g1 cat]

Sample

Al-SBA-15-10
Al-SBA-15-25
Al-SBA-15-40
Al-SBA-TMB(0.75)-10
Al-SBA-TMB(0.75)-25
Al-SBA-TMB(0.75)-40
FeW/Al-SBA-15-10
FeW/Al-SBA-15-25
FeW/Al-SBA-15-40
FeW/Al-SBA-TMB(0.75)-10
FeW/Al-SBA-TMB(0.75)-25
FeW/Al-SBA-TMB(0.75)-40

Fig. 9.

27

III and IIA (pH:3.11)

Silanol (pH:5.53)

Mixed group (IA) (pH:9.25)

1.14
0.66
2.27
0.51
1.36
1.23
1.57
1.21
1.82
2.57
0.54
0.67

0.91
1.55
23.54
1.04
3.16
11.14
3.21
5.63
5.88
7.77
2.75
3.27

2.35
0.94
5.47
1.65
8.99
3.97
5.29
25.55
10.29
20.54
3.21
4.12

Al MAS NMR spectra for Al modied supports.

related to the mixed group was only evidenced when the active
metals were incorporated (Fig. 8b). This result indicates that the
incorporation of FeW decreases the acid strength of the catalysts
(Table 2).
3.1.5. 27 Al MAS NMR solid measurements
Fig. 9 shows the 27 Al MAS NMR spectra for Al-SBA-15-x and
Al-SBA-TMB(0.75)-x. These materials exhibit two clear signals at a
chemical shift, = 0 ppm ascribed to AlO6 octahedral coordination
group, which is characteristic of the presence of extra structural
aluminum species and other signal at = 52.9 ppm corresponding
to AlO4 tetrahedral aluminum species [43].
The area of each Al peak was computed by a Gaussian integration
of the peaks (Section 2.4.4). In all cases, samples prepared with
a Si/Al molar ratio of 25 exhibited a greater incorporation of Al
within the lattice of SBA-15 supports, as evidenced by the intensity
of the signal corresponding to tetrahedral aluminum coordination
in Fig. 10. This result conrms the formation of Brnsted acid sites
for all catalysts as in agreement to proposed PAD results previously
commented (Fig. 8).
Octahedral aluminum species were found in higher proportion
on the Al-SBA-TMB(0.75)-x supports with a Si/Al molar ratio of 40
support (Fig. 10b). Such a change might be associated to the changes
in the morphology, which went from brous to spherical silica after
the addition of TMB. In this sense, the surface of the SBA-15 based
materials seems to be inuencing the kind of Al species introduced
to the framework of the material.

Fig. 10. Al species peak area for 27 Al MAS NMR spectra, (a) Al-SBA-15-x, (b) Al-SBATMB(0.75)-x, x indicates Si/Al molar ratio.

The post synthesis method of incorporation or chemical grafting proposed to incorporate on the silica framework aluminum
species was satisfactory to generate the acid strength distribution
to the material on the surface of the catalysts supported on these
modied silica.
These results are consistent with the PAD curves, which showed
that the materials with higher content of type III and IIA acid sites
were Al-SBA-15-25 and Al-SBA-TMB(0.75)-25. It is known that the
tetrahedral Al coordination is a precursor of Brnsted acid sites and
octahedral for Lewis acid sites [7,2730,44,45].

106

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

Fig. 11. Sulded catalysts tested selectivity at 7.3% of conversion.

Fig. 12. Sulded catalysts activity correlated to SBET , (a) FeW/Al-SBA-15-x, (b) FeW/Al-SBA-TMB(0.75)-x, x = Si/Al molar ratio.

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109


Table 3
Sulded catalysts initial rate constants.

107

4.2. Effect of acid properties in the HCK of phenanthrene

Sample

kPhen [mL g1 min1 ]

NiMo/-Al2 O3
FeW/-Al2 O3
FeW/Al-SBA-15-10
FeW/Al-SBA-15-25
FeW/Al(40)-SBA-15
FeW/Al-SBA-TMB(0.75)-10
FeW/Al-SBA-TMB(0.75)-25
FeW/Al-SBA-TMB(0.75)-40
Al-SBA-TMB(0.75)-10

2.16
0.18
0.21
0.52
0.69
0.84
0.43
0.32
0.12

3.2. Catalytic evaluation


3.2.1. Catalysts activity
Table 3 represents the change in the activity of the catalysts
evaluated. The commercial NiMo/-Al2 O3 catalyst exhibited the
highest initial rate constant (kPhen = 2.16 mL g1 min1 ). Among the
SBA-15 based catalysts prepared herein, the catalyst FeW/AlSBA-TMB(0.75)-10 was more active in comparison to the other
Al-modied TMB catalysts and Al-SBA-15 supported catalysts.
3.2.2. Catalysts selectivity
Fig. 11 depicts the iso-conversion selectivity towards condensable products. Selectivity is analyzed at a conversion of
7.3%. In particular the selectivity to 9,10-dihydrophenanthrene,
1,2,3,4-tetrahydrophenanthrene and trans-stilben is analyzed. The
commercial catalyst NiMo/-Al2 O3 as well as for FeW/-Al2 O3
were selective to hydrogenated (Fig. 13). FeW/Al-SBA-15-x and
FeW/Al-SBA-TMB(0.75)-x sulded catalysts were highly selective
to trans-stilben, which is a product of aromatic ring opening reactions. In general all FeW sulded catalysts showed no signicant
differences in selectivity.
4. Discussion
4.1. Effect of the textural properties in the HCK of phenanthrene
Fig. 12(a) show the catalytic activity of all SBA-15 based
catalysts, demonstrating the effect of specic surface area in
the catalytic activity of the sulded catalysts. The catalysts,
which showed the highest activity were: FeW/Al-SBA-15-40 and
FeW/Al-SBA-TMB(0.75)-10 sulded catalysts.
For both series of sulded catalysts, opposing trends in catalytic
activity by SBET were evidenced. Sulded catalysts supported on
Al-SBA-15-x exhibited that catalytic activity increased at the same
rate as SBET . Nevertheless, as indicated in Fig. 12(b), sulded catalysts supported on Al-SBA-TMB(0.75)-x did not correlate with the
increase in SBET . Thus, catalysts activity was not directly inuenced
by the porosity of this kind of TMB modied supports.
To provide a better explanation to this behavior wide angle XRD
patterns (Fig. 4) made it possible to establish that catalysts with
higher specic surface area, but higher pore volumes have enabled
the formation of larger particles of the metallic phase (Table 1).
Therefore, the catalytic activity with TMB modied catalysts was
diminished despite of their larger pore sizes.
On the other hand, SBET offers the required platform for homogeneous dispersion of metal nanoparticles. However, the effective
surface area required for metal dispersion suffered at the expense
of increasing pore diameter. The decrease in the mesostructured
ordering pattern of SBA-15 by the addition of TMB to expand its
pores in Al-SBA-TMB(0.75)-x supported catalysts as shown in Figs.
2(b) and 3 indicates less uniformity on pore size distribution and
pore structure, and thus decreasing SBET while widening pore size.

The mechanism of hydrocracking reactions is bi-functional [46],


which indicates that the occurrence of mild acid sites on the catalysts surface inuences the rate of reaction as the same selectivity
does. For that reason, the effect of Si/Al molar ratio on the performance of catalysts is elucidated in Fig. 13. For sulded catalysts
supported on Al-SBA-15-x, the highest Si/Al molar ratio leaded to
the highest activity. According with PAD analysis, overall acidity is
the highest for the highest Si/Al molar ratio (25, and 40).
For sulded catalysts supported on Al-SBA-TMB(0.75)-x, PAD
analysis given (Table 2) specied that when FeW oxidized species
were introduced on the supports, overall acidity was changed.
FeW/Al-SBA-TMB(0.75)-10 revealed the highest content of acid
sites. As a result, the lowest Si/Al molar ratio with TMB modied
sulded catalysts yielded in the highest catalytic activity results.
On the other hand, the inuence of acid strength is shown in
Fig. 14. The formation of ring opening compounds such as transstilben can be attributed to the effect of Al incorporation into the
lattice of SBA-15 based supports, as a result of the slight enhancement and tting in selectivity for the catalysts prepared with the
Si/Al molar ratio of 25, as compared to octahedral aluminum, which
revealed no correlation in sulded catalysts selectivity as shown in
Fig. 15. Therefore, it is established the effect of tetrahedral Al incorporation in the selectivity of phenanthrene hydrocracking reaction.
Tetrahedral aluminum species are precursors of Brnsted acid sites,
which in terms of acidity are stronger than Lewis acid sites related
to octahedral aluminum species.
A similar behavior was also evidenced by Dai et al., who tested
NiW/Al-SBA-15 catalysts in the hydrocracking of a VGO, demonstrating that Al-SBA-15 catalysts induced high selectivity to MD and
rupture compounds as a result of a higher tetrahedral aluminum
incorporation [27].
The previous results were directly correlated with the computed
values of 27 Al NMR spectra, which evidenced that at the highest
tetrahedral Al ions were incorporated to catalysts to Si/Al molar
ratio of 25 (Fig. 12). PAD deconvolution data as reported in Table 2,
also conrmed that the catalysts with higher content of type III acid
sites are the catalysts the highest Si/Al molar ratio. In that sense,
tetrahedral aluminum incorporation generates the strong acid sites
contribution to the overall acidity of the catalysts and supports.
Brnsted acidity is formed by this kind of Al incorporation, as it has
been reported by several authors, it inuences isomerization and
cracking reactions [2023,43,47,48].

Fig. 13. Sulded FeW/Al-SBA-15-x and FeW/Al-SBA-TMB(0.75)-x catalysts activity vs. Si/Al molar ratio, x indicates Si/Al molar ratio.

108

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

Fig. 14. Sulded catalysts selectivity correlated to AlO4 species, (a) FeW/Al-SBA-15-x, (b) FeW/Al-SBA-TMB(0.75)-x, x = Si/Al molar ratio.

Fig. 15. Sulded catalysts selectivity correlated to AlO6 species, (a) FeW/Al-SBA-15-x, (b) FeW/Al-SBA-TMB(0.75)-x, x = Si/Al molar ratio.

J.R. Restrepo-Garcia et al. / Applied Catalysis A: General 510 (2016) 98109

4.3. Effect of the FeW active phases in the HCK of phenanthrene


The effect of FeW active phase is evidenced by the changes
in catalytic activity when incorporated on Al-modied SBA-15
based catalysts. As shown in Table 3, catalytic activity with AlSBA-TMB(0.75)-10 support is 0.12 mL g1 min1 in comparison to
0.84 mL g1 min1 when FeW active phases are impregnated for
the same support. Therefore, as evidenced by several authors
[2326], a metallic phase is required to direct the reaction to the
desired products, and also to increase the activity and product yield.
In this case, the catalysts based on a FeW active phase exhibited a different product distribution as compared to conventional
sulde NiMo/-Al2 O3 catalyst. FeW catalysts supported on Al
incorporated silica were selective to ring opening compounds in
the hydrocracking of phenanthrene (trans-stilben). It is noteworthy that FeW/-Al2 O3 did not show the same product distribution,
conrming that Al-modied silica supports enhance the production
of such compounds, which is associated to the acid sites distribution
therein.
5. Conclusions
The synthesis parameters in SBA-15 mesostructured silica
preparation were modied. This modication led to get larger pore
diameter mesoporous silica by adding either a micellar expansion
agent such as TMB or an inorganic salt such NH4 F. The swelling
effect of TMB expanded micelles to form the pores to change morphology from a hexagonal array of tubular channels into spherical
nanoparticles, meanwhile adding small amounts of NH4 F yielded in
an amorphous silica material with higher pore diameters, but there
was a signicant loss in the specic surface area. Therefore, NH4 F
additions disfavored ordering at the initial synthesis temperature.
A series of Al-SBA-15 mesoporous materials were synthesized
by a post-synthesis aluminum grafting procedure. The incorporation of Al species led to an enhancement of the concentration of
Brnsted and Lewis acid sites, which was correlated to the formation of tetrahedral and octahedral aluminum surface species.
Therefore, generating stronger acid sites on the SBA-15 based materials.
FeW/Al-SBA based catalysts displayed high selectivity towards
the ring opening of phenanthrene due to their Brnsted acid sites.
It must be remarked that though the catalytic activity of the
FeW/Al-SBA based catalysts was lower than that of conventional
NiMo/Al2 O3 , the latter was only selective to partial hydrogenation
products. Therefore, FeW/Al-SBA based catalysts are promising
candidates for hydrocracking reactions.
Acknowledgements
This work was possible due to the nancial support of the
de catalizadores para
VIE-UIS in the frame of the project Diseno
hidrocraqueo de fracciones tipo gasleo y estudio del efecto de
molculas nitrogenadas, code 1329. Jonatan R. Restrepo thanks
to COLCIENCIAS for its Joven Investigador 2012 fellowship.
References
[1] J.L. Agudelo, B. Mezari, E.J.M. Hensen, S.A. Giraldo, L.J. Hoyos, Appl. Catal. A:
Gen. 488 (2014) 219230.
[2] S. Ahmed Ali, M. Elias Biswas, T. Yoneda, T. Miura, H. Hamid, E. Iwamatsu, H.
Al-Suaibi, in: H. Hideshi, O. Kiyoshi (Eds.), Studies in Surface Science and
Catalysis, Elsevier, 1999, pp. 407410.
[3] A.M. Alsobaai, R. Zakaria, B.H. Hameed, Fuel Process. Technol. 88 (2007)
921928.

109

[4] H. Ortiz-Moreno, J. Ramrez, F. Sanchez-Minero, R. Cuevas, J. Ancheyta, Fuel


130 (2014) 263272.
[5] J.P. Franck, J.F. Le page, in: T. Seiyama, K. Tanabe (Eds.), Studies in Surface
Science and Catalysis, Elsevier, 1981, pp. 792803.
[6] C.R. Lahiri, D. Biswas, Physica B+C 139140 (1986) 725728.
[7] X. Sun, H. Fan, J. Zhu, Shiyou Huagong/Petrochem. Technol. 42 (2013)
518523.
[8] P.E. Boahene, K.K. Soni, A.K. Dalai, J. Adjaye, Appl. Catal. A: Gen. 402 (2011)
3140.
[9] P.E. Boahene, K.K. Soni, A.K. Dalai, J. Adjaye, Catal. Today 207 (2012) 101111.
[10] K. Chandra Mouli, K. Soni, A. Dalai, J. Adjaye, Appl. Catal. A: Gen. 404 (2011)
2129.
[11] R.N. Widyaningrum, T.L. Church, M. Zhao, A.T. Harris, Int. J. Hydrog. Energy 37
(2012) 95909601.
[12] E. Byambajav, Appl. Catal. A: Gen. 252 (2003) 193204.
[13] B. Li, J. Xu, X. Li, J. Liu, S. Zuo, Z. Pan, Z. Wu, Mater. Res. Bull. 47 (2012)
11421148.
[14] X. Zhang, F. Zhang, X. Yan, Z. Zhang, F. Sun, Z. Wang, D. Zhao, J. Porous Mater.
15 (2007) 145150.
[15] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D.
Stucky, Science 279 (1998) 548552.

[16] D. Zhao, Y. Wan, in: H.v.B.A.C. Jir Cejka,


S. Ferdi (Eds.), Studies in Surface
Science and Catalysis, Elsevier, 2007, pp. 241III.
[17] M. Song, C. Zou, G. Niu, D. Zhao, Chin. J. Catal. 33 (2012) 140151.
[18] K. Waldron, Z. Wu, D. Zhao, X.D. Chen, C. Selomulya, Chem. Eng. Sci. 127
(2015) 276284.
[19] A.J.J. Koekkoek, J.A.R. van Veen, P.B. Gerrtisen, P. Giltay, P.C.M.M. Magusin,
E.J.M. Hensen, Microporous Mesoporous Mater. 151 (2012) 3443.
[20] S. Garg, K. Soni, G.M. Kumaran, M. Kumar, J.K. Gupta, L.D. Sharma, G.M. Dhar,
Catal. Today 130 (2008) 302308.
[21] M. Gmez-Cazalilla, A. Infantes-Molina, R. Moreno-Tost, P.J. Maireles-Torres,
J. Mrida-Robles, E. Rodrguez-Castelln, A. Jimnez-Lpez, Catal. Today 143
(2009) 137144.
[22] G. Muthu Kumaran, S. Garg, K. Soni, M. Kumar, L.D. Sharma, G. Murali Dhar,
K.S. Rama Rao, Appl. Catal. A: Gen. 305 (2006) 123129.
[23] T. Klimova, J. Reyes, O. Gutirrez, L. Lizama, Appl. Catal. A: Gen. 335 (2008)
159171.
[24] Z. Lei, L. Gao, H. Shui, W. Chen, Z. Wang, S. Ren, Fuel Process. Technol. 92
(2011) 20552060.
[25] C. Liang, M.-C. Wei, H.-H. Tseng, E.-C. Shu, Chem. Eng. J. 223 (2013) 785794.
[26] J.A. Mendoza-Nieto, I. Puente-Lee, C. Salcedo-Luna, T. Klimova, Fuel 100
(2012) 100109.
[27] Y. Dai, Y.S. Zhou, Q. Wei, Q.Y. Cui, Z. Qin, Ranliao Huaxue Xuebao/J. Fuel Chem.
Technol. 41 (2013) 15021506.
[28] M.G. Seo, D.W. Lee, K.Y. Lee, D.J. Moon, Fuel 143 (2015) 6371.
[29] S. Zeng, J. Blanchard, M. Breysse, Y. Shi, X. Shu, H. Nie, D. Li, Microporous
Mesoporous Mater. 85 (2005) 297304.
[30] J. Zhu, J. Wang, X. Sun, J. Yang, Petrol. Process. Petrochem. 43 (2012) 2832.
[31] Y. Villasana, F. Ruscio-Vanalesti, C. Pfaff, F.J. Mndez, M.. Luis-Luis, J.L. Brito,
Fuel 110 (2013) 259267.
[32] F. Rouquerol, J. Rouquerol, K. Sing, Adosrption by Powders and Porous Solids.
Principles, Methodology and Applications, 1999.
[33] C. Contescu, J. Jagiello, J.A. Schwarz, in: J.M.B.D.P.A.J.G. Poncelet, P. Grange
(Eds.), Studies in Surface Science and Catalysis, Elsevier, 1995, pp. 237252.
[34] C. Contescu, V.T. Popa, J.B. Miller, E.I. Ko, J.A. Schwarz, J. Catal. 157 (1995)
244258.
[35] H. Knzinger, P. Ratnasamy, Catal. Rev. 17 (1978) 3170.
[36] B.S. Gevert, J.-E. Otterstedt, Biomass 14 (1987) 173183.
[37] J. Rouquerol, D. Avnir, C.W. Fairbridge, D.H. Everett, J.H. Haynes, N. Pernicone,
J.D. Framsay, K. Sing, K.K. Unger, Pure Appl. Chem. 66 (1994) 17391758.
[38] R. Fazaeli, H. Aliyan, M.A. Ahmadi, S. Hashemian, Catal. Commun. 29 (2012)
4852.
[39] R. Gao, X. Yang, W.-L. Dai, Y. Le, H. Li, K. Fan, J. Catal. 256 (2008) 259267.
[40] P. Sharma, S.D. Park, K.T. Park, J.H. Park, C.Y. Jang, S.C. Nam, I.H. Baek, Powder
Technol. 233 (2013) 161168.
[41] A. Hassan, S. Ahmed, M.A. Ali, H. Hamid, T. Inui, Appl. Catal. A: Gen. 220
(2001) 5968.
[42] W.-H. Chen, Q. Zhao, H.-P. Lin, Y.-S. Yang, C.-Y. Mou, S.-B. Liu, Microporous
Mesoporous Mater. 66 (2003) 209218.
[43] G. Muthu Kumaran, S. Garg, K. Soni, M. Kumar, J.K. Gupta, L.D. Sharma, K.S.
Rama Rao, G. Murali Dhar, Microporous Mesoporous Mater. 114 (2008)
103109.
[44] B. Al Alwan, E. Sari, S.O. Salley, K.Y.S. Ng, Ind. Eng. Chem. Res. 53 (2014)
69236933.
[45] J. Blanchard, M. Breysse, K. Fajerwerg, C. Louis, C.E. Hdoire, A. Sampieri, S.
Zeng, G. Prot, H. Nie, D. Li, Stud. Surf. Sci. Catal. (2005) 15171524.
[46] J. Francis, E. Guillon, N. Bats, C. Pichon, A. Corma, L.J. Simon, Appl. Catal. A:
Gen. 409410 (2011) 140147.
[47] M. Gmez-Cazalilla, J.M. Mrida-Robles, A. Gurbani, E. Rodrguez-Castelln, A.
Jimnez-Lpez, J. Solid State Chem. 180 (2007) 11301140.
[48] T. Klimova, M. Caldern, J. Ramrez, Appl. Catal. A: Gen. 240 (2003) 2940.

Das könnte Ihnen auch gefallen