Sei sulla pagina 1di 7

Applied Clay Science 146 (2017) 432–438

Contents lists available at ScienceDirect

Applied Clay Science


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

Research paper

Ni-exchanged cationic clays as novel heterogeneous catalysts for selective MARK


ethylene oligomerization
Asma Aida,b, Radu Dorin Andreia, Samira Amokraneb, Claudia Cammaranoa, Djamel Niboub,
Vasile Huleaa,⁎
a
Institut Charles Gerhardt, UMR 5253, CNRS-UM2-ENSCM-UM1, Matériaux Avancés pour la Catalyse et la Santé, 8 rue de l'Ecole Normale, 34 296, Cedex 5,
Montpellier, France
b
Laboratoire de Technologie des Matériaux, Université des Sciences et de la Technologie Houari Boumediene, BP 32, El-Alia, Bab-Ezzouar, Alger, Algeria

A R T I C L E I N F O A B S T R A C T

Keywords: Montmorillonite rich clays, including bentonite, acid washed clay (K10) and Al-pillared clay were exchanged
Nickel with nickel in order to prepare novel heterogeneous catalysts for the ethylene oligomerization. The catalytic
Montmorillonite properties were examined in flow mode (p = 3.0 MPa, T = 150–350 °C), and the results were related to the
K10 textural and acidic characteristics of the catalysts. Due to higher surface area and mesoporous volume, as well as
Al-pillared clay
moderate acidity, Ni-exchanged K10 exhibited superior catalytic behaviour compared to others catalysts. Ni-K10
Ethylene
Oligomerization
was very active (TOF = 570 h− 1), stable (high conversion during 20 h on stream), and highly selective to linear
C4 and C6 olefins.

1. Introduction oligomerization catalyzed by the Ni-exchanged clays. The conversion of


ethylene into higher olefins by oligomerization reactions is of con-
Clays are abundant and low cost mineral resources of major im- siderable interest, and significant research efforts have been directed to
portance for a large variety of uses (Murray 1991; Wypych and the development of heterogeneous catalysts for these processes. Among
Satyanarayana 2004; Bergaya et al. 2006; Mukherjee 2013). Due to them, the Ni-exchanged inorganic porous materials are the most pro-
their specific properties, including the cation exchange, the swelling mising catalysts (Finiels et al. 2014). For these materials both the
ability and the thermal stability, clays are intensively explored as het- nature of the Ni sites and the texture of the catalyst are crucial variables
erogeneous catalysts for organic and inorganic reactions (Gil et al. affecting the catalytic activity and the stability. For example, the most
2000; Adam and McCabe, 2006; Dasgupta and Török 2008; active sites in this reaction are the nickel cations in exchange positions
Nagendrappa 2011; Zhou 2011; Daştan et al. 2012; Kumar et al. 2014). (Lacarriere et al. 2012; Finiels et al. 2014). In addition, the pore size of
For catalytic purpose, some key properties of clays, such as the specific the catalyst is very important. Thus, the Ni-exchanged zeolites, i.e. Ni-Y
surface area and the porosity, can be highly improved by treatment and Ni-MCM-22, suffered severe deactivation due to the blocking of
with a mineral acid solution (Kumar et al. 1995) or by intercalating their micropores with heavy products (Lallemand et al. 2006, 2008),
metal polycations between the layers (pillaring process) (Kloprogge while the Ni-exchanged catalysts with larger pores, i.e. Ni-AlMCM-41,
1998; De Stefanis and Tomlinson 2006). Although clays have been most Ni-AlSBA-15, and Ni-SiO2 exhibited very high productivity and stability
frequently used as H-form acid catalysts, the Me-exchanged clays in the ethylene conversion to C4-C12 olefins (Hulea and Fajula 2004;
(Me = Fe3 +, Cu2 +, Ti4 +, Al3 +, Zr4 +, Zn2 +, Ce3 +, Cr3 +, Sn2 +) Lallemand et al. 2009; Lacarriere et al. 2012; Andrei et al. 2015a,
proved to have interesting catalytic properties in various reactions, 2016).
including cycloaddition (Cabral et al. 1986), aldol condensation In order to develop new porous oligomerization catalysts, the re-
(Dumitriu et al. 1994), alkylation (Cseri et al. 1995), oxidation (Nikalje search was extended to the Ni-exchanged clay materials. For this pur-
and Sudalai 1999; Pillai and Sahle-Demessie 2003), hydroamination pose, montmorillonite rich clays with various chemical and textural
(Joseph et al. 2005), SCR of NOx (Dorado et al. 2006), acetylation properties, including an untreated bentonite, commercial acid treated
(Shimizu et al. 2008), acetalization (Thomas et al. 2011), hydroxyalk- K10 clay and an Al-pillared clay, were used as supports for preparing
ylation (Jha et al. 2013), etc. the catalysts. Montmorillonite, the representative member of the 2:1
This work relates to a novel application, namely the ethylene smectite family, is the most intensively explored clay mineral in


Corresponding author.
E-mail address: vasile.hulea@enscm.fr (V. Hulea).

http://dx.doi.org/10.1016/j.clay.2017.06.034
Received 20 December 2016; Received in revised form 26 June 2017; Accepted 28 June 2017
Available online 05 July 2017
0169-1317/ © 2017 Elsevier B.V. All rights reserved.
A. Aid et al. Applied Clay Science 146 (2017) 432–438

heterogeneous catalysis. When montmorillonite are washed with mi- Table 1


neral acids, the surface area increases notably, while the pore diameters Chemical composition (wt%) of Ni-based catalysts.
increase too, and the pores assume a three-dimensional form (Butruille
Sample Ni Si Al Mg Fe K Na Ti
and Pinnavaia 1992). Metal oxide pillared clays also possess very in-
teresting properties, such as large surface area, high pore volume and Ni-Bent 3.15 31.03 11.85 1.54 2.67 0.27 0.24 0.10
tunable pore size, and high thermal stability (Butruille and Pinnavaia Ni-K10 1.14 34.05 9.91 0.74 2.00 1.49 0.04 0.29
Ni-PILC 1.07 31.40 13.52 1.68 0.78 0.18 0.24 0.10
1992; Kloprogge 1998).
The Ni-containing clays were prepared by ionic exchange, and used
for the first time as catalysts in the ethylene oligomerization. The cat- NETZSCH TG 209C apparatus. The temperature program started with
alytic behaviors, evaluated in a flow mode reactor, were related to the an isothermal period of 5 min at 50 °C followed by a temperature ramp
chemical and textural properties of the catalysts. up to 900 °C at 10 °C min− 1 in synthetic air (20 mL min− 1). To eval-
uate the turnover frequencies (TOF, the number of molecules of ethy-
2. Experimental lene transformed by nickel site per hour) each Ni atom was considered
as a single accessible active site. Deactivation rates (in h− 1) were es-
2.1. Materials and catalysts timated from the parameter ((a0 − at)/(a0·t)) (where a0 is the initial
conversion and at is the conversion after t h on stream).
Ni-based catalysts were obtained by ion exchange of bentonite (Bent
- Sigma Aldrich), acid treated montmorillonite (K10 - Fluka) and
montmorillonite (Al) pillared clay (PILC - Aldrich), using Ni 3. Results and discussions
(NO3)2·6H2O (98%) (Alfa Aesar). Typically, 4 g of clay was contacted
for 24 h at 70 °C, under constant agitation, with 200 cm3 of 0.5 M 3.1. Material characteristics
aqueous solution of nickel nitrate, to obtain the catalytic materials
denoted as Ni-Bent, Ni-K10 and Ni-PILC. The solids were filtered, dried The chemical compositions of Ni-exchanged clays are listed in
in air at 80 °C for 12 h and then calcined at 550 °C, for 5 h, under a Table 1. Using the ionic exchange procedure described in the experi-
nitrogen flow. mental section, the amount of nickel incorporated in Ni-K10 and Ni-
PILC was about 1.1 wt%, whereas in Ni-Bent it was three times higher.
2.2. Characterization The Si/Al molar ratio was 2.6, 3.3 and 2.2 in Ni-Bent, Ni-K10 and Ni-
PILC, respectively.
The chemical composition of the materials was determined by ele- As previously reported by our group (Lallemand et al. 2009), the Ni
mental analysis using the X-Ray Fluorescence method. The phase sites in zeolites and mesoporous materials become active only after the
identification of the calcined samples was carried out by XRD on a high-temperature thermal treatment. For this reason, before char-
Bruker AXS D8 diffractometer, using Cu Kα radiation and a Ni filter. acterization and reaction, the clays used as catalysts in this work were
The samples were scanned from 2θ = 2 to 80° in steps of 0.02°. The calcined at 550 °C. In Fig. 1 are shown the XRD patterns of Ni-ex-
textural characterization of calcined samples was achieved by using a changed clays before any thermal treatment and after calcination. The
conventional gas adsorption/desorption method, at − 196 °C, on a non-calcined samples showed expected XRD patterns, typical for mon-
Micromeritics ASAP 2020 automatic analyzer. Prior to analysis, the tmorillonite with layered structures. Ni-Bent exhibited a basal distance
sample was treated under vacuum at 250 °C for 10 h. The specific (d001) of about 1.20 nm, while the basal distance of Ni-PILC was
surface area, SBET, was calculated in the domain of validity of the BET 1.71 nm. The d001 value of Ni-K10 was 1.44 nm, but the low intensity of
equation. The pore diameters have been evaluated from the nitrogen this signal indicates that the layered structure was disordered by the
sorption isotherms, using the Broekhoff and de Boer method. The acid treatment.
acidity of solids was measured by temperature programmed desorption All samples contain quartz as major impurity (2θ = 20.68 and
(TPD) using ammonia as probe. Prior to TPD experiments, the materials 26.5°). In addition, K10 sample contains illite as impurity (2θ = 8.8°).
were pre-treated in air flow at 550 °C for 5 h. Ammonia was adsorbed Calcination of the Ni-PILC and Ni-Bent samples only slightly changed
for 30 min at 100 °C and the physisorbed ammonia was removed by the XRD patterns, signifying that their structure are stable at high
evacuation of sample at 120 °C for 0.5 h, in a dry helium stream. The temperature. In contrast, the calcined Ni-K10 sample lost its layered
ammonia desorption was carried out in helium stream at a heating rate structure.
of 10 °C min− 1 up to 600 °C. The amount of desorbed ammonia was The N2 adsorption–desorption isotherms at − 196 °C of calcined Ni-
monitored with a TCD detector. X-ray photoelectron spectroscopy (XPS) exchanged clays are shown in Fig. 2. In the case of Ni-Bent, the hys-
measurements were performed on an ESCALAB 250 (Thermo Electron) teresis loop could be qualified as the H4 type, corresponding to the slit-
spectrometer equipped with an Al Kα source (1486.6 eV). like pores, present in a small amount. The isotherms of Ni-PILC and Ni-
K10 are intermediate between type I and type IV, with H3 type hys-
2.3. Catalytic oligomerization teresis, indicating a microporous-mesoporous structure. Based on ad-
sorption data, the specific surface areas, total pore volumes and average
The ethylene oligomerization was performed in flow mode in a diameter of the mesopores were measured and summarized in Table 2.
stainless steel fixed-bed reactor (i.d. 5 mm) using 1.0 g of catalyst Bentonite shows a surface area of 48 m2 g− 1, which is 3–4 times
(0.15–0.25 mm sieve fraction). The pressure was regulated via a back- lower than those of the modified clays. Significant differences also exist
pressure regulator. Before each test, the catalyst was activated in the between the pore volumes of the catalysts. Ni-K10 has a pore volume of
reactor at 550 °C under nitrogen flow for 8 h. Catalytic tests were 0.29 cm3 g− 1, which is twice or four times higher than that of Ni-PILC
performed sequentially at 150 °C, 200 °C, 250 °C, 300 °C and 350 °C, and Ni-Bent, respectively. The average diameter of mesopore was about
holding the reactor at each temperature for 1 h at 3.0 MPa of ethylene 5–7 nm for Ni-K10 and Ni-PILC (Fig. S1 in Supplementary data).
(without inert carrier gas), and mass hourly space velocity (WHSV) of Temperature-programmed desorption (TPD) of ammonia was car-
3.75 h− 1. The complete reactor effluent was analyzed online by gas ried out to measure the surface acidity of samples. The quantitative
chromatography (Varian CP-3800, FID), using a CP-PoraPLOT Q ca- evaluation (Table 2) indicates that the total number of acid sites in Ni-
pillary column (25 m, 0.53 mm, 20 μm) and an auto-sampling valve. Bent is significantly lower compared to Ni-K10 and Ni-PILC. The TPD
For the quantification of the products captured in the catalyst, the profiles plotted in Fig. 3 show that the Ni-exchanged clays have a large
spent catalyst (20 mg) was analyzed by thermogravimetry (TGA) with a proportion of acid sites with a weak strength (ammonia desorption

433
A. Aid et al. Applied Clay Science 146 (2017) 432–438

Fig. 1. X-ray diffraction patterns of Ni-K10, Ni-PILC and


Ni-Bent before (left) and after calcination at 550 °C (right).

Intensity (a.u.)
Intensity (a.u.)

Ni-K10 Ni-K10

Ni-PILC Ni-PILC

Ni-Bent
Ni-Bent

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
2 Theta (degrees) 2 Theta (degrees)

200 band centered at around 200–250 °C) and only a low amount of acid
sites with medium strength (T desorption > 350 °C).
180
Ni-K10 The TPD profiles of non-exchanged clays (Fig. S2 and Table S1 in
Volume adsorbed (cm3/g, STP)

160 the Supplementary data) also consisted of large ammonia desorption


140 band corresponding to the weak acid sites.
Moreover, the Ni-exchanged samples contain a higher overall
120 acidity than the original clays. This result is consistent with those
Ni-PILC
100 previously obtained with Ni-MCM-41 (Hulea and Fajula 2004) and Ni-Y
catalysts (Lallemand et al. 2006). During the ionic exchange, the Ni2 +
80
ions replaced most of the initial NH4+ ions, and after calcination, bi-
60 functional catalysts, having both Ni2 + and H+ sites, were obtained.
40 Ni-Bent Additionally, Lewis acid sites are present on the surface. For example,
using pyridine FT-IR, we found that, compared to H-Y zeolite, the Ni-Y
20 sample possessed a smaller fraction of Brønsted acid sites, but it showed
0 a higher concentration of Lewis acid sites. The excess of Lewis sites
0,0 0,2 0,4 0,6 0,8 1,0 were attributed to a Lewis type interaction of the pyridine with the
Relative Pressure (P/P0) Ni2 + ions.
To discriminate between the different types of nickel species present
Fig. 2. N2 adsorption-desorption isotherms at − 196 °C of the calcined Ni-K10, Ni-PILC in the catalysts, X-ray photoelectron spectroscopy was used. Fig. 4
and Ni-Bent samples. compares the Ni 2p core level spectra of Ni-K10 and Ni-PILC with that
of the pure nickel oxide (NiO). In agreement with the literature
Table 2 (Garbarino et al. 2016), the spectrum of NiO shows split both the main
Textural properties and acidity of catalysts. Ni 2p3/2 peak at 853.7, with a shoulder at 855.6 eV and their doublet Ni
Sample SBET Pore volume Mesopore Acidity
2p1/2 signal at 871.5 and 873.5 eV. In contrast, simpler spectra, con-
(m2 g− 1) (cm3 g− 1) diameter (nm) (mmolNH3 g− 1) sisting in a couple of two bands at 856.9 and 864.7 eV (2p3/2 and 2p1/2,
respectively) were obtained for both Ni-K10 and Ni-PILC catalysts.
Ni-K10 191 0.29 5–6 0.15 These values of the binding energies are similar to those previously
Ni-PILC 142 0.15 6–7 0.25
Ni-Bent 48 0.07 – 0.06
reported (Sohn et al. 2002) and are assigned to Ni2 + charge-compen-
sating cations. These results suggest that in both Ni-K10 and Ni-PILC
catalysts the Ni species are essentially isolated Ni2 + cations (no bulk
NiO species are present). It is important to note that the isolated Ni2 +
cations are considered as precursors for generating the active site in
oligomerization reactions (Lacarriere et al. 2012; Finiels et al. 2014).
Adsorbed ammonia, a.u.

Ni-PILC
853,7
855,6
Ni-K10 871,5
873,5
Intensity (a.u.)

NiO
Ni-BENT
856,9
Ni-K10 864,7

100 200 300 400 500


Temperature, °C Ni-PILC

Fig. 3. TPD profiles of Ni-exchanged clays.


885 880 875 870 865 860 855 850 845
Binding energy (eV)

Fig. 4. Ni 2p core level spectra of Ni-K10, Ni-PILC and NiO.

434
A. Aid et al. Applied Clay Science 146 (2017) 432–438

100 result indicates that, under the given conditions, the sites present on the
surface of this material (e.g. the acid sites and the species, such Fe, Mg,
K, Na, Ti) are not able to catalyze suitably the ethylene oligomerization.
80 The ethylene conversions obtained with the Ni-containing catalysts
Ethylene conversion, %

are plotted in Fig. 5.


In the presence of Ni-Bent the ethylene conversion was very low,
despite the high amount of nickel (3.15 wt% Ni in Ni-Bent, see Table 1).
60
Most likely, the Ni sites are not accessible for the reagent molecules
because of the thin layers distance or the small pore diameter. In con-
trast, Ni-PILC and Ni-K10 showed remarkable catalytic properties. In
40 the temperature range of 150–350 °C, over Ni-K10, the ethylene con-
version continuously increased, to reach 88% at 350 °C. In the presence
of Ni-PILC the ethylene conversion strongly increased up to 200–250 °C,
20 but after that, it reached a plateau at about 80%. Such a behaviour at
higher temperature could be attributed to a slightly deactivation suf-
fered by the catalyst during the catalytic test. This point will be dis-
0 cussed later.
150 200 250 300 350 Based on the results in Fig. 5, we found that at 250 °C the turnover
frequency was about 570 h− 1 for both Ni-PILC and Ni-K10 catalysts.
Temperature, °C
For comparison, the TOF value obtained at the same temperature over
Fig. 5. Effect of the reaction temperature on the ethylene conversion over Ni-K10 (□), Ni- Ni-AlSBA-15 (pore diameter of 9.5 nm) was 750 h− 1 (Andrei et al.
PILC (Δ) and Ni-Bent (◊). Conditions: WHSV = 3.75 h− 1, p = 3.0 MPa. 2015a). Note that this last result was the best one ever reported in the
literature with Ni-based porous catalysts without using alkylaluminium
3.2. Catalytic ethylene oligomerization cocatalysts.
Under the experimental conditions, for all catalysts, the main pro-
The catalytic behaviors of Ni-Bent, Ni-K10 and Ni-PILC were eval- ducts consisted of C4, C6, and C8 olefins. C4 fraction was exclusively
uated in flow mode, using a fixed-bed dynamic reactor, as described in linear, with an average composition of 18% 1-C4, 38% trans-2-C4 and
the experimental section. A blank test, using Ni free K10 as catalyst was 44% cis-2-C4. The C6 oligomers were mainly linear 2-hexenes (~ 75%),
also performed. With the Ni free sample the ethylene conversion varied while the C8 were mainly branched hydrocarbons. For industrial ap-
from 1 to 3% when the temperature increased from 150 to 350 °C. This plications, the challenge is to produce selectively alpha-olefins.

Fig. 6. Typical chromatogram of products obtained over Ni-K10


catalyst at 150 °C (a), 200 °C (b) and 250 °C (c).

435
A. Aid et al. Applied Clay Science 146 (2017) 432–438

Unfortunately, in the presence of Ni-based solid catalysts, the primary C8= C8=
1-olefins easily convert to 2-olefins (see below). However, the selective
CH3
formation of 2-butenes and 2-hexenes could be considered as a valuable
acid site Ni site
result. We reported in recent papers, that the ethylene can be directly H 3C H2C CH2
converted into propylene by one-pot catalytic cascade reactions, using
two heterogeneous catalysts (Andrei et al. 2015b, 2016). In a single CH3 H2C CH2
flow reactor ethylene is first selectively dimerized/isomerized over Ni- 2 H2C CH2
H 2C
(CH2)3 CH3
Ni site
based catalyst to form 2-butenes, which react then with the excess of H 2C Ni site

ethylene in a metathesis reaction catalyzed by Mo-based oxides, to acid site


acid site
produce propylene. This technology is already used to face the higher
growth rate in the demand for propylene compared to ethylene.
CH3 (CH2)2 - CH3
In the oligomerization process performed on Ni-clays, besides bu-
tenes, hexenes and octenes, olefins with odd carbon number (C3, C5 H 3C H 3C
acid site
and C7) and alkanes were also formed in process. Some traces of C10+
hydrocarbons were present among the products. Heavy carbonaceous
species (denoted here below as “coke”) were formed on the catalyst C10=
surface.
The amount and the nature of the products strongly depended on
the reaction temperature, as shown from the typical chromatograms in cracking/hydrogen trasfer
C6, C8, C10+ lower alkanes/olefins + coke
Fig. 6, which contains the hydrocarbon distributions obtained on Ni- acid site

K10 at 150, 200 and 250 °C. Scheme 1. Main reaction pathways in the ethylene oligomerization process.
At low temperature, C4 is the major product (> 90 mol%), while at
higher temperature, the oligomerization is also directed towards the 90
formation of C6 (16–18 mol%) and C8 olefins (6–8 mol%)(Table 3). The
formation of alkanes and C3, C5, C7 olefins indicates that, beside 80
ethylene oligomerization, reactions such as hydrogen transfer and hy-
Ethylene conversion, %

drocarbon cracking occurred, particularly at higher temperatures. 70


These experimental results are consistent with those previously
obtained with other Ni-based catalysts (Hulea and Fajula 2004; 60
Lacarriere et al. 2012; Lallemand et al. 2006, 2008) and confirm the
existence of different types of reactions, as shown in Scheme 1. The 50
initial Ni-catalyzed reaction is the oligomerization of ethylene yielding
1-butene. Nickel also acts as catalytic site for further oligomerization 40
reactions involving butene-ethylene and hexene-ethylene to form C6
30
and C8 olefins. When the desorbed 1-olefins migrate on an acid site,
they are easily converted to internal double-bond olefins. Note that
20
double-bond shift in alkenes is an extremely easy reaction, exhibiting 0 5 10 15 20
turnover frequencies up to 107 s− 1 in heterogeneous acid catalysis
Time on stream, h
(Haag et al. 1984). C4 and C6 olefins can be consumed through di-
merization reactions (on acid sites), yielding C8 and C10 + oligomers. Fig. 7. Ethylene conversion vs. time on stream for Ni-K10 (□) and Ni-PILC (Δ).
These heavy products are involved in acid-catalyzed cracking reactions, Conditions: T = 250 °C, WHSV = 3.75 h− 1, p = 3.0 MPa.

to form lower hydrocarbons, particularly at high temperature. The


hydrogen transfer reactions, typically associated with the cracking re- deactivation rates are completely different. Ni-PILC shows a relatively
actions (Abbot and Wojciechowski 1987), also occur on the acid sites. short lifetime (1/2 of the initial conversion after 8.5 h), with a deacti-
They are responsible for the formation of saturated hydrocarbons vation rate of 3.1 10− 2 h− 1. In contrast, for Ni-K10, the activity de-
(mainly CH4, C2H6, C3H8 and C4H10) and coke. clined smoothly during 20 h, with an average deactivation rate of 1.2
To gain insight into the catalyst stability (deactivation is a major 10− 2 h− 1. This result is better than those obtained with Ni-exchanged
drawback in heterogeneous processes), long-term experiments, over dealuminated Y zeolite (with an average deactivation rate of 1.9
20 h on-stream were carried out at 250 °C, 3.0 MPa of ethylene and 10− 2 h− 1 for 20 h on stream) (Heveling et al. 1988), but lower than
WHSV = 3.75 h− 1 on Ni-K10 and Ni-PILC fresh catalysts. In Fig. 7 are those obtained with Ni-AlSBA-15 mesoporous materials (average de-
plotted the conversions of ethylene vs. time on stream (TOS) for both activation rate of 1.6 10− 3 h− 1) (Andrei et al. 2015a).
catalysts. The catalysts exhibited high initial activity, but the TG measurements under oxidizing conditions were carried out on

Table 3
Effect of the temperature on the product distribution.a

Catalyst T, °C CH4 C2H6 C3H6 C3H8 C4H8 C4H10 C5b C6b C7b C8b

Ni-K10 150 0 0.4 0.1 0 93.2 0.4 1.2 3.9 0.3 0.5
250 < 0.1 3.5 0.2 < 0.1 71.7 0.1 1.2 16.3 0.7 6.2
300 0.1 5.9 0.7 < 0.1 64.2 < 0.1 1.7 17.8 1.1 8.4
350 0.5 9.8 1.6 0.2 56.9 0.1 3.1 17.7 2.2 7.9
Ni-PILC 150 0 0.4 0.2 0.2 90.1 < 0.1 0.4 6.7 0.5 1.4
250 < 0.1 1.8 < 0.1 0.3 71.6 < 0.1 0.6 17.9 0.8 6.9
300 0.1 3.7 < 0.1 0.6 69.8 < 0.1 0.8 16.4 1.2 7.3

a
Mol %.
b
Olefins + alkanes.

436
A. Aid et al. Applied Clay Science 146 (2017) 432–438

Table 4 successfully prepared from commercial clays by ion exchange with


Effect of the TOS on the product distribution. nickel. The prepared Ni-K10 and Ni-PILC catalysts (1.1 wt% Ni), with
both micropores and mesopores, and high specific surface areas, re-
TOS (h) Catalyst Product distribution, mol% Others
vealed promising properties for ethylene oligomerization. Using a fixed-
C2H6 C3H6 C4H8 C5 C6 C7 C8 bed flow reactor, the catalysts showed good activity and stability. The
reaction was highly selective to olefins with an even number of carbon
1.1 Ni-K10 2.9 0.3 74.5 0.6 14.8 0.2 6.4 0.3
atoms. However, the catalytic behaviour (activity, selectivity, stability)
7.5 2.1 0.1 74.7 0.6 15.3 0.2 6.8 0.2
15.1 1.5 0.1 76.8 0.3 14.3 0.1 6.5 0.4 of the catalysts strongly depended on their textures and acidic prop-
19.6 0.3 0.1 84.8 0.1 9.5 0.1 4.1 0.4 erties. Thus, Ni-exchanged K10, which combines high pore accessibility
1.1 Ni-PILC 3.5 0.3 76.6 0.9 14.1 0.7 3.6 0.3 and lower acid site density, exhibited superior catalytic behaviour
4.1 1.6 0.1 78.6 0.2 14.7 0.8 3.7 0.3 compared to Ni-PILC sample. It proved to be a versatile, robust and
7.7 1.0 0.1 83.6 0.1 10.7 0.5 3.8 0.2
efficient catalyst for the selective oligomerization of ethylene.

Table 5
Acknowledgement
Effect of the TOS on the distribution of butenes.
Asma Aid had a financial support from the French and Algerian
TOS (h) Catalyst Butene distribution, mol% Ministries of Foreign Affairs (PROFAS B + program).
1-C4H8 2-trans-C4H8 2-cis-C4H8
Appendix A. Supplementary data
1.1 Ni-K10 17 38 45
7.5 17 39 44 Supplementary data to this article can be found online at http://dx.
15.1 19 39 42
doi.org/10.1016/j.clay.2017.06.034.
19.6 21 40 39
1.1 Ni-PILC 18 37 45
4.1 27 35 38 References
7.7 34 34 32
Abbot, J., Wojciechowski, B.W., 1987. Hydrogen transfer reactions in the catalytic
cracking of paraffins. J. Catal. 107, 451–462.
the Ni-K10 and Ni-PILC spent catalysts in order to evaluate the amount Adam, J.M., McCabe, R.W., 2006. Clay minerals as catalysts. In: Bergaya, F., Theng,
B.K.G., Lagaly, G. (Eds.), Handbook of Clay Science. Elsevier, Amsterdam, pp.
of products confined into the pores after 20 h on stream. The total
541–581.
weight loss was only 3.5% for the spent Ni-Bent, proving the absence of Andrei, R.D., Popa, M.I., Fajula, F., Hulea, V., 2015a. Heterogeneous oligomerization of
the reaction on this catalyst. In contrast, the weight loss was 10% and ethylene over highly active and stable Ni-AlSBA-15 mesoporous catalysts. J. Catal.
17% for Ni-K10 and Ni-PILC, respectively. In both cases, the weight loss 323, 76–84.
Andrei, R.D., Popa, M.I., Fajula, F., Cammarano, C., Hulea, V., Al Khudhair, A.,
consisted in two fractions: 1–2% in the range of 50–300 °C (light and Bouchmella, K., Mutin, P.H., Hulea, V., 2015b. Ethylene to propylene by one-pot
middle oligomers) and a main loss in the range of 300–650 °C, corre- catalytic cascade reactions. ACS Catal. 5, 2774–2777.
sponding to the combustion of heavy products/coke species. The Andrei, R.D., Mureseanu, M., Popa, M.I., Cammarano, C., Fajula, F., Hulea, V., 2015c. Ni-
exchanged AlSBA-15 mesoporous materials as outstanding catalysts for ethylene
average rate of the coke formation (in mg of coke per g of catalyst and oligomerization. Eur. Phys. J. Special Topics 224, 1831–1841.
hour on stream) was 8.5 for Ni-PILC and only 5.0 for Ni-K10. This Andrei, R.D., Popa, M.I., Cammarano, C., Hulea, V., 2016. Nickel and molybdenum
difference may explain why Ni-K10 (having higher surface area and containing mesoporous catalysts for ethylene oligomerization and metathesis. New J.
Chem. 40, 4146–4152.
mesopore volume) showed a lower deactivation rate. Indeed, as re- Bergaya, F., Theng, B.K.G., Lagaly, G. (Eds.), 2006. Handbook of Clay Science. Elsevier,
cently shown for the Ni-AlSBA-15 catalyst (Lin et al. 2014; Andrei et al. Amsterdam.
2015a, 2015c), the mesopores facilitate the diffusion of the oligomer- Butruille, J.R., Pinnavaia, T.J., 1992. Propene alkylation of liquid phase biphenyl cata-
lyzed by alumina pillared clay catalysts. Catal. Today 14, 141–154.
ization products, particularly the bulkier ones, resulting in a lower
Cabral, J., Laszlo, P., Montaufier, M.T., 1986. Catalysis of the cyclohexadienone-phenol
deactivation rate of the catalysts. On the other side, the high density of rearrangement by a Lewis-acidic clay system. Tetrahedron Lett. 27, 2627–2630.
the acid sites in Ni-PILC could be responsible for the formation of the Cseri, T., Bekassy, S., Figueras, F., Rizner, S., 1995. Benzylation of aromatics on ion-
exchanged clays. J. Mol. Catal. A 98, 101–107.
heavy products and the fast deactivation of this catalyst. A similar effect
Dasgupta, S., Török, B., 2008. Application of clay catalysts in organic synthesis. A review.
of the high acidity on the catalyst deactivation was also reported for the Org. Prep. Proced. Int. 40, 1–65.
ethylene oligomerization catalyzed by the Ni-zeolites (Lallemand et al. Daştan, A., Kulkarni, A., Török, B., 2012. Environmentally benign synthesis of hetero-
2006). cyclic compounds by combined microwave-assisted heterogeneous catalytic ap-
proaches. Green Chem. 14, 17–37.
As shown in Table 4, for both Ni-K10 and Ni-PILC catalysts, the De Stefanis, A., Tomlinson, A.A.G., 2006. Towards designing pillared clays for catalysis.
product distribution depended on the time on stream. At high values of Catal. Today 114, 126–141.
TOS, when the catalysts were partially deactivated, the ratio C4/ Dorado, F., de Lucas, A., Garcia, P.B., Romero, A., 2006. Preparation of Cu-ion-exchanged
Fe-PILCs for the SCR of NO by propene. Appl. Catal. B. 65, 175–184.
(C6 + C8) increased. Additionally, the ratio 1-butene/2-butenes in- Dumitriu, E., Balba, N., Lupascu, M., Azzouz, A., Hulea, V., Cirje, G., Nibou, D., 1994.
creased (Table 5). These results suggest that the deactivated sites are Synthesis of acroleinby vapor-phase condensation of formaldehyde and acetaldehyde
mainly the acid sites, which are responsible for the double bond iso- over synthetic and natural zeolites. J. Catal. 147, 133–139.
Finiels, A., Fajula, F., Hulea, V., 2014. Nickel-based solid catalysts for ethylene oligo-
merization and co-oligomerization reactions. merization – a review. Catal. Sci. Technol. 4, 2412–2426.
The efficiency of a regeneration process was evaluated on the used Garbarino, G., Riani, P., Infantes-Molina, A., Rodríguez-Castellón, E., Busca, G., 2016. On
Ni-K10 and Ni-PILC catalysts. It consisted of calcination at 550 °C for the detectability limits of nickel species on NiO/γ-Al2O3 catalytic materials. Appl.
Catal. A 525, 180–189.
8 h with a mixture air-nitrogen (25v/75v). The regenerated catalyst
Gil, A., Gandia, L.M., Vicente, M.A., 2000. Recent advances in the synthesis and catalytic
exhibited catalytic properties (activity and product distribution) in all applications of pillared clays. Catal. Rev. Sci. Eng. 42, 145–212.
points analogous to those of the original catalyst. As example, Fig. S3 Haag, W.O., Lago, R.M., Weisz, P.B., 1984. The active site of acidic aluminosilicate cat-
alysts. Nature 309, 589–591.
(in the Supplementary data) compares the ethylene conversions for the
Heveling, J., Van Der Beek, A., De Pender, M., 1988. Oligomerization of ethene over
fresh and regenerated Ni-K10. nickel-exchanged zeolite Y into a diesel-range product. Appl. Catal. A. 42, 325–336.
Hulea, V., Fajula, F., 2004. Ni-exchanged AlMCM-41 – an efficient bi-functional catalyst
for the ethylene oligomerization. J. Catal. 225, 213–222.
4. Conclusions Jha, A., Garade, A.C., Shirai, M., Rode, C.V., 2013. Metal cation-exchanged montmor-
illonite clay as catalysts for hydroxyalkylation reaction. Appl. Clay Sci. 74, 141–146.
Joseph, T., Shanbhag, G.V., Halligudi, S.B., 2005. Copper(II) ion-exchanged
This study showed that porous inexpensive catalysts can be

437
A. Aid et al. Applied Clay Science 146 (2017) 432–438

montmorillonite as catalyst for the direct addition of NeH bond to CC triple bond. J. Murray, H.H., 1991. Overview - clay mineral applications. Appl. Clay Sci. 5 (5–6),
Mol. Catal. A Chem. 236, 139–144. 379–395.
Kloprogge, J.T., 1998. Synthesis of smectites and porous pillared clay catalysts: a review. Nagendrappa, G., 2011. Organic synthesis using clay and clay-supported catalysts. Appl.
J. Porous. Mater. 5, 5–41. Clay Sci. 53, 106–138.
Kumar, P., Jasra, R.V., Bhat, T.S.G., 1995. Evolution of porosity and surface acidity in Nikalje, M.D., Sudalai, A., 1999. Catalytic selective oxidation of alkyl arenes to aryl tert-
montmorillonite clay on acid activation. Ind. Eng. Chem. Res. 34, 1440–1448. butyl peroxides with TBHP over Ru-exchanged montmorillonite K 10. Tetrahedron
Kumar, B.S., Dhakshinamoorthy, A., Pitchumani, K., 2014. K10 montmorillononite clays 55, 5903–5908.
as environmentally bening catalysts for organic reactions. Catal. Sci. Technol. 4, Pillai, U.R., Sahle-Demessie, E., 2003. Oxidation of alcohols over Fe3 +/montmorillonite-
2378–2396. K10 using hydrogen peroxide. Appl. Catal. A 245, 103–109.
Lacarriere, A., Robin, J., Swierczynski, D., Finiels, A., Fajula, F., Luck, F., Hulea, V., 2012. Shimizu, K., Higuchi, T., Takasugi, E., Hatamachi, T., Kodama, T., Satsuma, A., 2008.
Distillate-range products from non-oil based sources by catalytic cascade reactions. Characterization of Lewis acidity of cation-exchanged montmorillonite K10 clay as
ChemSusChem 5, 1787–1792. effective heterogeneous catalyst for acetylation of alcohol. J. Mol. Catal. A Chem.
Lallemand, M., Finiels, A., Fajula, F., Hulea, V., 2006. Catalytic oligomerization of 284, 89–96.
ethylene over Ni-containing dealuminated Y zeolites. Appl. Catal. A 301, 196–201. Sohn, J.R., Park, W.C., Kim, H.W., 2002. Characterization of nickel sulfate supported on
Lallemand, M., Rusu, O.A., Dumitriu, E., Finiels, A., Fajula, F., Hulea, V., 2008. NiMCM- γ-Al2O3 for ethylene dimerization and its relationship to acidic properties. J. Catal.
36 and NiMCM-22 catalysts for the ethylene oligomerization. Effect of zeolite texture 209, 69–74.
and nickel cations/acid sites ratio. Appl. Catal. A 338, 37–43. Thomas, B., Ramu, V.G., Gopinath, S., George, J., Kurian, M., Laurent, G., Drisko, G.L.,
Lallemand, M., Finiels, A., Fajula, F., Hulea, V., 2009. Nature of the active sites in Sugunan, S., 2011. Catalytic acetalization of carbonyl compounds over cation (Ce3 +,
ethylene oligomerization catalyzed by Ni-containing molecular sieves: chemical and Fe3 + and Al3 +) exchanged montmorillonites and Ce3 +-exchanged Y zeolites. Appl.
IR spectral investigation. J. Phys. Chem. C 113, 20360–20364. Clay Sci. 53, 227–235.
Lin, S., Shi, L., Zhang, H., Zhang, N., Yi, X., Zheng, A., Li, X., 2014. Tuning the pore Wypych, F., Satyanarayana, K.G. (Eds.), 2004. Clay Surfaces: Fundamentals and
structure of plug-containing Al-SBA-15 by post-treatment and its selectivity for C16 Applications. Elsevier, Amsterdam.
olefin in ethylene oligomerization. Microporous Mesoporous Mater. 184, 151–161. Zhou, C.H., 2011. An overview on strategies towards clay-based designer catalysts for
Mukherjee, S., 2013. The Science of Clays: Applications in Industry, Engineering, and green and sustainable catalysis. Appl. Clay Sci. 53, 87–96.
Environment. Springer.

438

Potrebbero piacerti anche