Вы находитесь на странице: 1из 15

Reac Kinet Mech Cat (2015) 114:295309

DOI 10.1007/s11144-014-0790-3

Effects of the activation method on the performance


of base metal catalysts prepared by wet impregnation
for toluene hydrogenation in liquid phase

Raphael Soeiro Suppino Richard Landers

Antonio Jose Gomez Cobo

Received: 9 June 2014 / Accepted: 23 September 2014 / Published online: 1 October 2014
Akademiai Kiado, Budapest, Hungary 2014

Abstract The present work aims to study the effects of the activation method on the
performance of base metal catalysts for toluene hydrogenation in liquid phase. For
this, catalysts of Fe, Co and Ni supported on c-Al2O3 were prepared by wet
impregnation from chlorinated precursors and reduced by formaldehyde. The pre-
pared catalysts were activated ex situ at 773 K or in situ at 523 K both under H2 and
characterized by N2 physisorption, SEM ? EDX, TEM, XPS and TPR techniques.
Catalytic tests were conducted in a slurry Parr reactor at 373 K under H2 pressure of
5 MPa. The results indicate the formation of metal hydroxides during the catalysts
preparation, which are not reduced by the formaldehyde. The ex situ activation seems
able to reduce both hydroxides of Fe and Ni, but there is no evidence of reduction in
the case of Co hydroxide. Co and Ni catalysts present higher activities when not
activated, whereas the in situ activation increases the activity of the Fe catalyst.

Keywords Base metals  Alumina  Wet impregnation  Toluene hydrogenation

Introduction

Supported metal catalysts have been extensively used in hydrogenation reactions.


Although noble metal catalysts are known for their higher activity and selectivity in

R. S. Suppino (&)  A. J. G. Cobo


Laboratory for Development of Catalytic Processes, Department of Chemical Systems Engineering,
School of Chemical Engineering, University of Campinas, Cidade Universitaria Zeferino Vaz,
Campinas, SP CEP 13083-852, Brazil
e-mail: rsuppino@gmail.com

R. Landers
Laboratory of Surface Physics, Department of Applied Physics, Gleb Wataghin Institute of
Physics, University of Campinas, Campinas, SP CEP 13083-859, Brazil

123
296 Reac Kinet Mech Cat (2015) 114:295309

hydrogenation reactions [1, 2], base metals are also of great interest mainly due to their
lower cost [14]. In addition, base metals such as Ni have presented reasonable resistance
to sulfur poisoning, which is very important in hydrodearomatization processes [1].
Among the base metals employed in hydrogenation reactions, Ni is the dominant
on an industrial scale and also in the research literature [28]. Co-based catalysts
have been employed in numerous studies of hydrogenation reactions [9, 10],
particularly in FischerTropsch reactions, given the intrinsic ability of Co to
hydrogenate dissociated carbon species and promote chain growth [11]. In turn, Fe
catalysts are less employed in hydrogenation studies [12].
Some authors have reported the enhancement of base metal catalytic perfor-
mances by the addition of a noble metal. The hydrogenation of aromatic compounds
in the presence of PtNi [13, 14], PtCo [13], PdNi [15] and Ru-base metals (such
as Fe, Co, Ni) [16] was assessed by previous works. In general, the results indicate
an improvement on both activity and selectivity of these catalysts with the addition
of the noble metals.
Several materials, such as SiO2, Al2O3 and zeolites, may be used as support for
base metal catalysts [3, 9, 10, 17] and the nature of the support may affect the
properties of the active phase. The metal interacts with the support and this
interaction has influence on both the dispersion and reducibility of the metal species
[3]. Due to its high surface area and thermal stability, c-Al2O3 has been widely used
in chemical processes, notably in association with Ni. Therefore, Ni/Al2O3 catalysts
are among the most employed on petrochemical and fine chemical synthesis [4].
Different factors may have influence on the hydrogenation performance of nickel
catalysts, especially with respect to the dispersion and reducibility of the metal.
These two properties are usually contradictory and are difficult to improve
simultaneously, hence a balance of them tend to optimize the active surface area of
Ni/Al2O3 catalysts [4].
Catalysts destined to aromatics hydrogenation may be prepared by many methods,
such as solgel [3], co-precipitation [4], incipient [1821] and wet impregnation [10,
2226]. According to Savva et al. [3], the solgel method ensures a better compromise
between the dispersion of nickel phase and its interaction with the support surface than
co-precipitation and incipient impregnation. However, Ni loadings higher than
15 wt% were not achieved by the solgel method, as the gelation was difficult and the
initially formed gel, if any, was not stable.
The impregnation methods are the most commonly employed, mainly due to the
simplicity of the procedures [27]. In a recent review, Maki-Arvela and Murzin [28]
stated that usually the metal dispersions are lower in catalysts prepared by incipient
impregnation than with wet impregnation. Bu et al. [23] and Suppino et al. [26]
demonstrated that wet impregnation leads to the formation of smaller metal particles
in comparison with incipient impregnation for supported Ru catalysts. The activity
of these solids on the partial hydrogenation of benzene were also superior, as well as
its selectivity to cyclohexene [23, 26].
It is important to emphasize that very few studies regarding the preparation of
base metal catalysts by wet impregnation are found in the literature [1, 10], while
this method is widely employed to prepare noble metal catalysts [22, 23, 26]
destined to hydrogenation of aromatics.

123
Reac Kinet Mech Cat (2015) 114:295309 297

The activation procedure on catalysts prepared by wet impregnation may vary


from liquid-phase reduction with NaBH4 [22] or formaldehyde [26] under mild
conditions to high temperature H2 flow ex situ reduction [3, 10, 23]. It is noteworthy
that depending on the metal nature, extreme reduction conditions may cause
sintering leading to a decrease on metal dispersion and catalytic activity [29].
The hydrogenation of aromatics is a reaction of great interest due to
environmental aspects and the wide range of industrial processes involving this
reaction. It is an important route to obtain several chemical intermediates as well as
it eliminates toxic aromatics compounds present on fuel [17, 30, 31]. Recently,
Alhumaidan et al. [32] proposed the use of the catalytic hydrogenation cycle
toluene-methylcyclohexane-toluene as hydrogen storage for fuel cells. Additionally,
the toluene hydrogenation is often used as a probe reaction to test the performance
of metal catalysts since it is, along with the benzene hydrogenation, taken as
insensitive with respect to surface structures [4].
Considering the exposed, the present work aims to study the effects of the
activation method on the performance of base metal (Fe, Co, Ni) catalysts supported
on alumina and prepared by wet impregnation for the toluene hydrogenation in the
liquid phase.

Experimental

Catalysts preparations

Gamma-alumina (c-Al2O3) of commercial grade (Alfa Aesar, 99.9 %) with mean


particle diameter of 58 lm was used as received as catalyst support. The specific
surface area of the support was determined by N2 physisorption, as detailed
hereafter.
The catalysts were prepared in order to obtain a metal fraction of 5 wt%, from
the precursors FeCl36H2O, NiCl26H2O (both from Aldrich Chemical Co.) and
CoCl26H2O (Merck), all with 99.9 wt% of purity.
In the wet impregnation method, deionized water was added to the support
resulting in a suspension continuously agitated at room temperature by a magnetic
stirrer. An aqueous solution of the precursor salt (Fe, Co or Ni) was then slowly
added to this suspension. Afterwards, the resultant suspension was heated to 353 K
and then its pH was adjusted to 10 by adding an aqueous solution of NaOH. This
adjustment is important in order to favor the approach of metal cations towards the
support, since the isoelectric point of c-Al2O3 is known to be around the pH of 7 and
8 [33]. Formaldehyde (Merck, 37 wt%) was used as a reducing agent. The
suspension was then filtrated in a Buchner funnel and the remaining solid was
washed with deionized water in order to remove residual chlorine, sodium and
formaldehyde. The solids were washed until a test with AgNO3 did not reveal
residual chloride and a flame test did not indicate the presence of sodium. The
catalysts were dried at 358 K for 24 h. All solids prepared by this method do not
present any abbreviation in its denominations.

123
298 Reac Kinet Mech Cat (2015) 114:295309

Part of the solids already reduced by formaldehyde were submitted to a


subsequent ex situ treatment under a 40 mL/min flow of H2 from the room
temperature until 773 K, remaining at this temperature for 3 h. The catalysts that
received this treatment have the E abbreviation on its denominations.
A second part of the catalysts reduced by formaldehyde were submitted to in situ
treatment under 3 MPa of H2 at 523 K for 1 h. The solids that received this
treatment present the I abbreviation. Ni/Al2O3 was not submitted to this treatment
given the little influence it showed on Co and Fe reduction as discussed ahead.

Support and catalysts characterizations

The specific surface area of the solids was determined through N2 physisorption
(BET method). A sample of 1.00 g of each solid was previously dried at 473 K
under vacuum and the physisorption was conducted at 77 K in a Tristar
Micromeritics ASAP 2010 instrument.
Scanning electron microscopy coupled with spectrometric X-ray analysis
(SEM ? EDX) was used with the principal aim of evaluating semi-quantitatively
the chemical composition of the catalysts. The analyses were conducted in a LEO
440i Leica instrument. Before insertion in the SEM, all samples were covered with a
fine layer of gold atoms using a 3 mA current for 180 s in order to obtain a gold film
thickness of 92 A .
TEM analyses were carried out on a Libra 120 Zeiss microscope with Cantega
2 k/Olympus CCD camera and iTEM data acquisition platform. The samples were
gently grinded and then dispersed in water. The dispersion was placed on ultrasound
for 10 min and then left to rest for another 10 min. A drop of the solution was
placed on a 300 mesh cupper grid coated with parlodium and carbon. The grids were
dried at ambient temperature and examined at 80 kV using the energy filter at zero
loss, 25, 30 or 50 eV. The energy positions 25 and 50 eV correspond to the first and
second plasmon.
X-ray photoelectron spectroscopy (XPS) analyses were conducted in order to
study the chemical composition on the surface of the catalysts. A spherical analyzer
VSWHA-100 with aluminum anode (Al Ka, hv = 1,486.6 eV) was used. The
pressure during the analyses was lower than 2 9 10-6 Pa. To correct bonding
energies, the line Al 2p with bonding energy of 74.0 eV was used as reference.
The formation of active phases in the catalysts was studied through temperature
programmed reduction (TPR). A Micromeritics Auto Chem 2910 instrument was
used to obtain the TPR profiles. In these analyses, a sample of 50 mg of each solid
was heated at 10 K/min from 298 to 1,073 K under 50 mL/min flow of a 10 % H2 in
N2 mixture.

Catalytic tests

The catalytic tests in the hydrogenation of toluene (Sigma Aldrich, 99 %) were


performed in a slurry Parr reactor at 373 K, under 5 MPa of H2 pressure and at a
stirring speed of 1,000 rpm.

123
Reac Kinet Mech Cat (2015) 114:295309 299

These tests were conducted with 300 mg of the catalyst, 25 mL of toluene,


30 mL of distilled water and 5 mL of n-heptane (Sigma Aldrich, 99 %) used as an
internal standard for the chromatographic method. The presence of water is
paramount in order to obtain partially hydrogenated products [34]. n-Heptane was
considered to be a compound suitable for the analytical purposes, because it is very
soluble in the organic phase (where the reaction products are) and easily separated
from the reaction products during the chromatographic analysis, besides being
chemically inert under the reaction conditions.
Samples of the organic phase were collected during the reaction and analyzed on
a HP 5890 series II gas chromatograph equipped with a flame ionization detector
and a 0.25 mm 9 25 m capillary column with dimethylsyloxane phase.
In order to determine the initial reaction rate, the instantaneous reaction rate was
calculated for each experimental point (concentration of toluene versus reaction
time). Through a linear fit of these rates, preferably for the values close to the
beginning of the reaction, an extrapolation of these values was made to zero
reaction time.

Results and discussion

Textural properties of the solids

As seen in Table 1, there is an increase of the specific surface areas of the catalysts
in comparison to the support (97 m2/g). This effect is considered to be related to the
formation of hydroxide species of the metals by the addition of NaOH during the
wet impregnation procedure [25], according to Eq. 1, exemplified by NiCl2.
2NaOH NiCl2 ! NiOH2 2NaCl 1

The increase of the specific surface area on catalysts prepared by wet


impregnation has been reported by Suppino et al. [26] for supported Ru catalysts.
These authors have also observed that catalysts that contained ruthenium hydroxide
are more easily reduced than those with chlorine species.
According to Table 1, the increase on the specific surface area depends on the
nature of the metal and is more pronounced on Ni/Al2O3. Transmission electron
microscopy of this catalyst (Fig. 1a) has shown the existence of lamellar structures
that were not found on the other solids (Fig. 1b, for example). It is therefore

Table 1 Specific surface area


Solid Sg (m2/g) Mass fraction (%) by EDX
(Sg) and elementary chemical
composition of solids Al O Cl Metal

Al2O3 97 44 56
Fe/Al2O3 109 42 54 0.0 4.2
Co/Al2O3 99 37 57 0.2 5.5
Ni/Al2O3 116 41 55 0.0 4.1

123
300 Reac Kinet Mech Cat (2015) 114:295309

Fig. 1 TEM images: a Ni/Al2O3 and b Fe/Al2O3, both reduced only by formaldehyde; c Fe/Al2O3
activated ex situ under H2 flow and d Fe/Al2O3 activated in situ under H2 pressure

possible that the Ni(OH)2 presents a different morphology than Co or Fe hydroxides


and this could be responsible for the distinct increase on the specific surface area of
the Ni/Al2O3 solid.

Elementary chemical composition of the solids

The elementary chemical composition of the solids, obtained through EDX analysis,
is also presented in Table 1.
Although this method is considered semi-quantitative, the results indicate that the
metal load on all catalysts are very close to the nominal amount (5 wt%). As stated
before, the impregnation pH has direct influence on the metal fixation. With this
result, one can infer that the chosen pH value of 10 led to a proper fixation of all
base metals in this study.

123
Reac Kinet Mech Cat (2015) 114:295309 301

It can also be seen in Table 1 that residual chlorine was only detected on the Co/
Al2O3 catalyst. This result suggests that not all CoCl2 have been converted to
Co(OH)2 during the preparation. In turn, the amount of chlorine is rather small,
around 10 % of the nominal quantity present on the precursor.
The XPS technique allowed the evaluation of the effect of subsequent H2
reduction on the elimination of residual chlorine. The atomic ratios obtained from
this analysis are shown in Table 2 in which the catalysts reduced only with
formaldehyde present no abbreviation, whereas the E abbreviation indicates that the
solid was submitted to a subsequent ex situ reduction. In turn, the I abbreviation
indicates that the solid was reduced in situ, besides its previous reduction with
formaldehyde.
Comparing the results of Table 2 and those obtained via EDX (Table 1), one can
observe that the XPS analysis reveals the presence of chlorine (Cl/Metal ratio) on
catalysts reduced by formaldehyde, in which the EDX analysis indicate the total
absence of this compound. Such differences might be due to a lower penetration of
XPS on the solids, which allows a better resolution of the surface where the active
metals and chlorine are probably concentrated.
Even though chlorine is detected, the Cl/Metal ratios indicate that the wet
impregnation method leads to an efficient elimination of most of the chlorine,
probably due to hydroxide formation. Indeed, whilst the theoretical Cl/Metal ratio
for Ni and Co precursors is 2.0 and for Fe is 3.0, a maximum of 0.10 Cl per metal
atom is observed. It is noteworthy that the subsequent reduction ex situ with H2 flow
decreases even more the amount of chlorine on the catalysts, especially on Fe/
Al2O3-E. This solid presents ten times less chlorine than the one reduced only by
formaldehyde (Fe/Al2O3).
The analysis of the binding energies of the metals was conducted with the XPS
technique and is also presented in Table 2. With these results, the probable species
present on each catalyst were identified based on XPS pattern data.
It is worth mentioning the possible presence of base metal hydroxide on the
surface of every catalyst reduced by formaldehyde. The existence of hydroxides is
consistent with the addition of NaOH with the purpose of adjusting the
impregnation pH (Eq. 1). Additionally, one can observe that the effects of the
activation treatments are, in general, strongly influenced by the nature of the metal.
The binding energy obtained for the iron catalyst (710 eV) suggest that, given a
0.5 eV range, Fe(OH)3 (709.5 eV) is not reduced with the addition of formaldehyde
(Fe/Al2O3 catalyst). As a matter of fact, the same binding energy may also indicate
the presence of FeCl3 (709.9 eV), according to the energy patterns. Indeed, as
previously discussed, chlorine was found on the surface of this solid. However, its
amount suggests that only a small quantity of the Fe may be found in the chloride
form. In turn, the catalyst submitted to ex situ H2 reduction presented a completely
different binding energy for the Fe species (706.8 eV), in this case related to its
metallic form. Therefore, the results suggest that only under high temperature and
H2 flow one can obtain Fe0. The mild conditions of liquid phase reduction with
formaldehyde are not suitable for this purpose.
Cobalt presented the very same binding energy, independent of the activation
method, and probably related to Co(OH)2. This result indicates that, despite of the

123
302

123
Table 2 Mean particle diameter of the metal (dp); XPS analysis on chlorine and surface metal species; H2 consumption and peak temperature on the TPR profiles of the
studied catalysts
Catalyst TEM analysis XPS analysis TPR analysis
a b
dp (nm) r (nm) Atomic ratio Cl/Metal B. E. (eV) Probable surface species T (K) H2 consumption (lmolH2 /mgmetal)

Per peak Total

Fe/Al2O3 4.4 1.5 0.10 710 FeCl3; Fe(OH)3 617 2.7 11


711 8.2
Fe/Al2O3-E 8.4 1.9 0.01 706.8 Fe0 594 7.0 7.0
Fe/Al2O3-I 5.7 1.0 n.a. n.a. n.a. 636 3.7 15
746 11
Co/Al2O3 3.1 0.9 0.08 781.5 Co(OH)2 609 3.1 9.0
718 6.2
Co/Al2O3-E 7.7 2.2 0.10 781.4 Co(OH)2 673 8.1 8.1
Co/Al2O3-I n.a. n.a. n.a. n.a. n.a. 649 1.3 8.5
723 7.2
Ni/Al2O3 7.0 0.7 0.03 856.1 Ni(OH)2 662 11 11
Ni/Al2O3-E n.a. n.a. 0.03 855.2 NiO 520 7.5 7.5

r standard deviation of the particles diameters, T temperature, n.a. sample not submitted to this analysis
a
B. E. binding energies for the metals Fe 2p3/2, Co 2p3/2, Ni 2p3/2
b
Binding energies patterns obtained from www.lasurface.com
Reac Kinet Mech Cat (2015) 114:295309
Reac Kinet Mech Cat (2015) 114:295309 303

formation of cobalt hydroxide, the reduction of this specie may be difficult, since
not even with the elevated temperature (773 K) of ex situ treatment Co0 was
formed. The Co binding energy identified in these catalysts (781.5 eV) also
indicates that the chlorine observed via EDX and XPS may be adsorbed on the
support, in opposition of CoCl2 (782.9 eV).
The binding energy obtained for the Ni/Al2O3 catalyst (856.1 eV) suggests that
the metal species found on this solid may be Ni(OH)2 (856 eV). Hence, the addition
of formaldehyde does not lead to the formation of Ni0. According to the results,
nickel hydroxide is reduced when the solid is submitted to ex situ H2 reduction,
since the binding energy found in this case (855.2 eV) is related to NiO (855.3 eV).
The nickel oxide may be evidence that the metal was reduced and suffered an
oxidation when exposed to atmospheric air during its manipulation.

Particle size

The Fe, Co and Ni/Al2O3 catalysts were submitted to TEM analysis with the
purpose of studying the effects of the activation methods on the particle size of the
metals. The results are shown in Table 2.
It is important to emphasize that under the conditions employed in this analysis,
one could not differentiate the particles in elementary state from those in chloride,
hydroxide or oxide forms. As seen via the XPS technique, the base metals might not
be completely reduced and therefore the particle diameter measured with TEM
cannot be directly related to metallic particles or metallic dispersion.
The reduction treatment led to an increase in Fe particles compared to the
catalyst reduced only with formaldehyde (Figs. 1b1d). The ex situ reduction, for
instance, led to the formation of Fe particles almost twice the size of the ones in Fe/
Al2O3 catalyst. The standard deviation of the diameters is higher for the solid
reduced ex situ, suggesting a heterogeneous distribution of the particles. This result
may be due to a coalescence of the metal particles, as suggested by Pattabiraman
et al. [29]. Despite the high Tamman temperature of Fe (1,043 K), the ex situ
reduction was conducted at a temperature, which could favor the mobility of iron
particles. The in situ treatment, conducted at lower temperature also led to an
increase on Fe particles diameters, but it was slighter than on the solid activated ex
situ.
It is possible that the mean particle diameter is also related to the metal species
on each catalyst surface. Therefore, the increase on the metal particle size could be
due to the reduction of Fe(OH)3 to Fe0. This result suggests that it is very difficult to
obtain metallic Fe particles without sacrificing some of its dispersion, which is in
agreement with previous findings reported in literature [4].
Similarly, for Co catalysts, an increase in the particle size was obtained
with ex situ reduction. The standard deviation of the particle sizes suggest a more
heterogeneous distribution, in comparison with the solid reduced only by
formaldehyde. The Co(OH)2 species was found on both Co/Al2O3 and Co/Al2O3-
E catalysts, which could reinforce the coalescence hypothesis.
Ni catalyst reduced with H2 was not submitted to TEM analysis. The solid
reduced only with formaldehyde (Ni/Al2O3) presented Ni species with both

123
304 Reac Kinet Mech Cat (2015) 114:295309

spherical and lamellar form (Fig. 1a), indicating a different morphology for
Ni(OH)2, in comparison to the Fe and Co. The mean particle diameter of Ni was
calculated with the few nearly spherical particles found in the analysis and the
standard deviation shows a narrow distribution of the Ni particles sizes.

Active phase formation

Table 2 shows the H2 consumption (per peak and total) as well as the temperatures
of the peaks in TPR profiles for the Fe/Al2O3, Co/Al2O3 and Ni/Al2O3 catalysts
reduced by formaldehyde and activated with H2 ex situ (E abbreviation) and in situ
(I abbreviation).
The presence of two peaks on the Fe/Al2O3 catalyst is consistent with the FeCl3
and Fe(OH)3 species identified with XPS. Since the H2 consumption for this catalyst
was 15 lmol of H2/mgFe and the theoretical consumption for the complete reduction
of the metal precursor (FeCl3) is 27 lmol of H2/mgFe, the TPR analysis suggests
that a small part of the metal may have been reduced with the addition of the
formaldehyde, although the XPS technique did not reveal the presence of Fe0.
The catalyst submitted to a subsequent reduction with H2 in situ (Fe/Al2O3-I)
presents similar behavior in comparison with the catalyst reduced only with
formaldehyde, both in temperature peaks and H2 consumption (ca. 13 2 lmol of
H2/mgFe). In addition, the H2 consumption at higher temperature suggests that the
reduction of FeCl3 requires more drastic treatments to be achieved. Indeed, in the
TPR analysis of Fe/Al2O3-E, one can observe a single peak at a much lower
temperature, which is consistent with XPS results that revealed the presence of Fe0
on this solid. The lower H2 consumption of this solid (7.0 lmol of H2/mgFe) may be
related to a partial oxidation of the Fe specie due to the exposure of the catalyst to
atmospheric air during its manipulation.
The in situ reduction had little influence on the formation of the active phase for
Co catalysts. Comparing the TPR profiles of Co catalysts in Fig. 2 and the H2
consumption in Table 2, the resemblance of the profiles of Co/Al2O3 and Co/Al2O3-
I catalysts is remarkable. Both of them presented two peaks, which may be related
to the interaction of Co with the support surface, as suggested by Jacobs et al. [35].
The reduction peaks of these solids are practically at the same temperature and
present close H2 consumptions (ca. 8.8 2 lmol of H2/mgCo), which are lower
than the stoichiometric value for the complete reduction of CoCl2 (17 lmol of H2/
mgCo). This result suggests that the addition of formaldehyde may have led to a
slight reduction of the metal, although the energy signature of metallic Co was not
identified with the XPS analysis.
Although a single peak was observed for the Co/Al2O3-E solid, the H2
consumption (8.0 lmol of H2/mgCo) is close to the other Co catalysts, suggesting
that the ex situ treatment could not completely reduce the Co(OH)2 present on this
solid. Nevertheless, evidence suggests that the subsequent ex situ treatment may
contribute by decreasing the temperature required for this reduction, as seen on the
TPR profile of Co/Al2O3-E and in Table 2.
In contrast to what was obtained on TPR profiles for Fe and Co catalysts, Ni/
Al2O3 presented a single peak, independent of the activation method. However, the

123
Reac Kinet Mech Cat (2015) 114:295309 305

(a)

H2 consumption (a.u.)

(b)

(c)

300 400 500 600 700 800 900 1000 1100


Temperature(K)

Fig. 2 TPR profiles of Co/Al2O3 catalysts: a reduced only by formaldehyde, b activated ex situ under H2
flow and c activated in situ under H2 pressure

ex situ reduction led to a more easily reducible solid (Ni/Al2O3-E), as suggested by


the lower temperature in which the H2 consumption peak was found.
The H2 consumption of the Ni/Al2O3 solid (11 lmol of H2/mgNi) suggests that
formaldehyde was unable to reduce the Ni(OH)2 significantly, since the stoichi-
ometric H2 consumption for the complete reduction of the Ni precursor is 17 lmol
of H2/mgNi. In turn, the H2 consumption of the solid reduced ex situ (Ni/Al2O3-E)
was much lower (7.5 lmol of H2/mgNi), indicating that the reduction treatment
under H2 flow led to the formation of Ni0. However, probably due to the exposure of
this solid to the air, part of the metal was oxidized, as previously shown by the XPS
analysis.

Catalytic performance for toluene hydrogenation

In order to evaluate the effects of the activation method on the performance of the
Fe, Co and Ni catalysts, these solids were tested in toluene hydrogenation. Table 3
presents the values of initial reaction rate (r0) and the conversion of toluene after
360 min of reaction (X).
From the analysis of Table 3, one may observe that the subsequent activation
treatments led to an increase in the activity of Fe catalysts, which may be related to
the species of Fe found on the surfaces of these solids. On the Fe/Al2O3
catalyst, reduced only by formaldehyde, the XPS analysis indicated the presence of
FeCl3, which may hinder the adsorption of the reagents, thereby decreasing the
catalytic activity. In contrast, on the Fe/Al2O3-E solid the XPS and TPR analyses
suggest that most of the Fe was in metallic form. It is probable that the existence of
Fe0 in the catalyst reduced ex situ increases the number of active sites, and
consequently the activity of this solid. In turn, the Fe/Al2O3-I catalyst presented a

123
306 Reac Kinet Mech Cat (2015) 114:295309

Table 3 Catalytic performance a


Catalyst r0 X (%)
on the toluene hydrogenation
Fe/Al2O3 1.3 2.3
Fe/Al2O3-E 2.0 3.4
Fe/Al2O3-I 2.7 3.4
Co/Al2O3 4.4 2.7
Co/Al2O3-E 1.3 3.5
X conversion of toluene after Co/Al2O3-I 1.7 2.8
360 min of reaction Ni/Al2O3 1.7 2.4
a
r0 in (mmoltolueneL-1 min-1 Ni/Al2O3-E 1.3 3.3
g-1
cat ): initial reaction rate

slightly higher initial activity, in comparison to the solid reduced ex situ. These
results indicate that the activation of Fe/Al2O3 with H2 and elevated temperature is
paramount to obtain more active solids. Moreover, the in situ treatment had little
effect on the Fe particle size, whereas the ex situ activation led to larger particles of
this metal.
Regarding the active forms of the metals, Taimoor and Pitault [36] have
presented a kinetic study of the gas phase hydrogenation of toluene with Pt
catalysts, in which they propose that both sites metallic and cationic (metalsupport
interaction) have an important role in this reaction. According to the authors, H2
would be adsorbed over the metallic sites, whereas toluene would be adsorbed over
cationic sites. Moreover, in a previous study, Mazzieri et al. [37] also considered
that the ratio between metallic and cationic Ru sites have influence on the selectivity
and activity of catalysts employed on hydrogenation of aromatics. In the present
work, catalytic activity has been observed with solids in which metallic species
could not be detected (Fe/Al2O3, for instance) as well as with solids that are almost
completely reduced (Fe/Al2O3-E). Therefore, the authors agree with Taimoor and
Pitault [36] and Mazzieri et al. [37], in the sense that both sites M0 and Md? are the
active forms of the metals in this study.
The conversion of toluene after 360 min followed the same behavior as the
activity, indicating that the Fe catalysts submitted to subsequent activation
treatments, especially in situ, led to more active and stable solids, not susceptible
to deactivation.
Co solids showed different catalytic behaviors depending on the reduction
treatment, as seen in Fig. 3. The subsequent reduction of Co catalysts with H2 led to
an initial activity comparable to Fe/Al2O3 (Table 3) and much lower in comparison
to the Co/Al2O3, solid reduced only by formaldehyde. As seen in the TPR analyses,
the reduction of Co species tends to be difficult even with higher temperatures
treatments (such as ex situ). In addition, an increase on the Co particle size was
observed with the reduction under H2 flow, which could be related to the lower
activity of this solid.
The higher activity among Co catalysts was observed for the solid reduced only
by formaldehyde. The performance of this solid is probably related to the intrinsic
catalytic properties of Co [11], associated to effects induced by the wet

123
Reac Kinet Mech Cat (2015) 114:295309 307

4,0
Co/Al2O3
3,5 Co/Al2O3-E
Co/Al2O3-I

Conversion of toluene (%)


3,0

2,5

2,0

1,5

1,0

0,5

0,0
0 50 100 150 200 250 300 350 400
Reaction time (min)

Fig. 3 Influence of the activation method on the performance of cobalt catalysts: reduced only by
formaldehyde (Co/Al2O3), activated ex situ under H2 flow (Co/Al2O3-E) and activated in situ under H2
pressure (Co/Al2O3-I). Reaction conditions 300 mg of the catalyst; 25 mL of toluene, 30 mL of distilled
water and 5 mL of n-heptane; reaction temperature 373 K, pressure 5 MPa of H2, stirring rate 1,000 rpm

impregnation procedure employed in the preparation of the catalysts. Nevertheless,


this high activity is mainly observed in the first hours of the reaction. After 300 min,
it seems that the catalyst suffers a deactivation and thus no toluene conversion
higher than 2.7 % could be achieved.
The deactivation of catalysts on liquid phase hydrogenation reactions has been
studied, but there is still no consensus about its causes. Lylykangas et al. [38]
studied Ni, Co and Pt catalysts employed in the hydrogenation of isooctenes in
liquid phase. The authors observed a deactivation of these solids, which was
attributed to the formation of carbonaceous species.
In turn, Taimoor and Pitault [36] reported that one of the most probable causes of
catalyst deactivation on toluene hydrogenation is the occupancy of metallic sites by
toluene molecules in detriment of H2 adsorption. According to theses authors, the
reaction occurs when the H2 molecules are adsorbed on metallic sites, whereas
toluene is adsorbed on metal-interaction sites. The determining step of the reaction
would be the transfer of the hydrogen from metallic to the metal-interaction sites,
occupied by toluene and where the reaction takes place.
Since the catalytic tests of the present study were conducted at a relatively low
temperature (373 K), the occurrence of classic mechanisms of deactivation in the
gas phase, such as sintering or coking, seems to be unlikely. Therefore, it is
suggested that the deactivation observed in the case of the Co/Al2O3 catalyst is
probably due to a competitive adsorption between H2 and toluene over the metal, as
proposed by Taimoor and Pitault [36].
Ni catalysts presented a similar behavior compared to Co/Al2O3 solids with
regard to the reduction treatment, since the initial reaction rate of Ni/Al2O3 was
higher than Ni/Al2O3-E. In this case, the higher activity of the first catalyst may be
related to the presence of Ni(OH)2, instead of the NiO present on the latter. It is

123
308 Reac Kinet Mech Cat (2015) 114:295309

possible that the Ni hydroxide is more easily activated on reaction conditions, than
the oxidized specie, which could lead to the formation of the active phase on the
beginning of the reaction.
Additionally, as previously discussed for Co, the toluene conversion after
360 min for the Ni/Al2O3 catalyst is lower than what was obtained for Ni/Al2O3-E.
Although the deactivation of the Ni/Al2O3 catalyst is not as strong as the Co/Al2O3,
in both cases the activity loss may be due to an increase of toluene adsorption over
these metals, according to the competitive adsorption model (H2 versus toluene)
proposed by Taimoor and Pitault [36].
Finally, it is noteworthy that despite the presence of water, no intermediate
reaction product, such as methylcyclohexene, was identified in any of the catalytic
tests conducted in this study. Therefore it can be inferred that the studied base
metals catalysts, regardless of the metal nature or activation method, are not
selective for partial hydrogenation products.

Conclusions

The present work studied the effects of the activation method on the performance of
base metal catalysts supported on c-Al2O3 for toluene hydrogenation in liquid
phase.
During the catalyst preparation by wet impregnation from chlorinated precursors,
metal hydroxides are formed, which are not reduced by the formaldehyde.
An ex situ activation, conducted at 773 K under H2 flow, seems able to reduce
both hydroxides of Fe and Ni, but there is no evidence of reduction in the case of Co
hydroxide.
Co and Ni catalysts present higher activities when not activated, whereas an
in situ activation at 573 K under H2 increases the activity of the Fe catalyst.

Acknowledgments To the National Counsel of Technological and Scientific Development (CNPq)


for the financial support and research scholarship granted to Raphael Soeiro Suppino.

References

1. Barrio V, Arias P, Cambra J, Guemez M, Pawelec B, Fierro J (2003) Appl Catal A-Gen 242(1):1730
2. Masalska A (2005) Appl Catal A-Gen 294(2):260272
3. Savva PG, Goundani K, Vakros J, Bourikas K, Fountzoula C, Vattis D, Lycourghiotis A, Kordulis C
(2008) Appl Catal B-Environ 79(3):199207
4. Hu S, Xue M, Chen H, Sun Y, Shen J (2011) Chin J Catal 32(68):917925
5. Toppinen S, Rantakyla T-K, Salml T, Aittamaa J (1997) Catal Today 38(1):2330
6. Lindfors LP, Salmi T (1993) Ind Eng Chem Res 32(1):3442
7. Rautanen PA, Aittamaa JR, Krause AOI (2000) Ind Eng Chem Res 39(11):40324039
8. Lylykangas MS, Rautanen PA, Krause AOI (2002) Ind Eng Chem Res 41(23):56325639
9. Backman LB, Rautiainen A, Krause AOI, Lindblad M (1998) Catal Today 43(12):1119
10. Szegedi A , Popova M, Mavrodinova V, Minchev C (2008) Appl Catal A-Gen 338(1):4451
11. Bianchi CL (2001) Catal Lett 76(34):155159
12. Yoon KJ, Vannice MA (1983) J Catal 82(2):457468
13. Lu S, Lonergan WW, Bosco JP, Wang S, Zhu Y, Xie Y, Chen JG (2008) J Catal 259(2):260268
14. Lonergan WW, Vlachos DG, Chen JG (2010) J Catal 271(2):239250

123
Reac Kinet Mech Cat (2015) 114:295309 309

15. Castano P, Pawelec B, Fierro J, Arandes J, Bilbao J (2007) Fuel 86(15):22622274


16. Sun H-J, Pan Y-J, Jiang H-B, Li S-H, Zhang Y-X, Liu S-C, Liu Z-Y (2013) Appl Catal A-Gen
464:19
17. Castano P, Pawelec B, Fierro J, Arandes J, Bilbao J (2006) Appl Catal A-Gen 315:101113
18. Mazzieri V, Coloma-Pascual F, Gonzalez M, LArgentie`re P, Fgoli N (2002) React Kinet Catal Lett
76(1):5359
19. Spinace EV, Vaz JM (2003) Catal Commun 4(3):9196
20. Wang J, Huang L, Li Q (1998) Appl Catal A-Gen 175(12):191199
21. Loiha S, Fottinger K, Zorn K, Klysubun W, Rupprechter G, Wittayakun J (2009) J Ind Eng Chem
15(6):819823
22. Fan G-Y, Jiang W-D, Wang J-B, Li R-X, Chen H, Li X-J (2008) Catal Commun 10(1):98102
23. Bu J, Pei Y, Guo P, Qiao M, Yan S, Fan K (2007) Stud Surf Sci Catal 165:769772
24. Liu J-L, Zhu L-J, Pei Y, Zhuang J-H, Li H, Li H-X, Qiao M-H, Fan K-N (2009) Appl Catal A-Gen
353(2):282287
25. Kawi S, Liu SY, Shen SC (2001) Catal Today 68(13):237244
26. Suppino RS, Landers R, Cobo AJG (2013) Appl Catal A-Gen 452:916
27. Jiao L, Regalbuto JR (2008) J Catal 260(2):329341
28. Maki-Arvela P, Murzin DY (2013) Appl Catal A-Gen 451:251281
29. Pattabiraman R (1997) Appl Catal A-Gen 153(1):920
30. Sato K, Aoki M, Noyori R (1998) Science 281(5383):16461647
31. Dietzsch E, Claus P, Honicke D (2000) Top Catal 10(12):99106
32. Alhumaidan F, Cresswell D, Garforth A (2011) Energ Fuel 25(10):42174234
33. Kosmulski M (2009) Surface charging and points of zero charge. CRC Press, USA
34. Johnson MM, Nowack GP (1975) J Catal 38(13):518521
35. Jacobs G, Das TK, Patterson PM, Li J, Sanchez L, Davis BH (2003) Appl Catal A-Gen
247(2):335343
36. Taimoor AA, Pitault I (2011) Reac Kinet Mech Cat 102(2):263282
37. Mazzieri VA, LArgentie`re PC, Coloma-Pascual F, Figoli NS (2003) Ind Eng Chem Res
42(11):22692272
38. Lylykangas MS, Rautanen PA, Krause AOI (2004) Ind Eng Chem Res 43(7):16411648

123

Вам также может понравиться