Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Energy & Fuels 1997, 11, 1225-1231 1225

Catalytic Conversion of Polyolefins into Liquid Fuels


over MCM-41: Comparison with ZSM-5 and Amorphous
SiO2-Al2O3
J. Aguado,* J. L. Sotelo, D. P. Serrano, J. A. Calles, and J. M. Escola
Chemical Engineering Department, Faculty of Chemistry, Complutense University of Madrid,
28040 Madrid, Spain

Received April 8, 1997X

The catalytic degradation of both low- and high-density polyethylene (LDPE and HDPE) and
polypropylene (PP) has been investigated using MCM-41, a mesoporous aluminosilicate recently
discovered, as catalyst. The results obtained have been compared to those of ZSM-5 zeolite and
amorphous silica-alumina. For all the studied plastics, MCM-41 has been found more active
than the amorphous SiO2-Al2O3, as a consequence of the higher surface area and the uniform
mesoporosity present in the former. Compared to ZSM-5, MCM-41 exhibits a lower activity for
the degradation of linear and low branched polymers (HDPE and LDPE, respectively), which
can be related to the higher strength of the zeolite acid sites. However, the opposite is observed
for the cracking of highly substituted plastics such as PP due to the severe steric hindrances
these molecules encounter to enter into the narrow pores of the zeolite, as confirmed by molecular
simulation measurements. Moreover, for the cracking of LDPE, HDPE, and PP, the selectivities
toward hydrocarbons in the range of gasolines and middle distillates obtained over MCM-41 are
clearly higher than those of ZSM-5. Therefore, MCM-41 is a catalyst potentially interesting for
the conversion of polyolefinic plastic wastes into liquid fuels.

Introduction recovered by hydrolysis, methanolysis, and glycolysis.


On the contrary, addition polymers such as polyolefins
The output of plastic materials has continuously
need stronger thermal or catalytic treatments to be
increased during the past decades, leading to a parallel
converted back into hydrocarbon mixtures useful as
rise in the generation of plastic wastes. The public
fuels or petrochemical feedstocks.7
concern over the environmental impact caused by
polymeric wastes, and the economic interest for a Thermal degradation of polyolefins is usually a low
profitable use of such abundant source of chemicals and selective process, leading to a wide distribution of waxy
energy, are the main driving forces of the current products8 that must be upgraded by a subsequent
research activities aimed at the development of plastic catalytic reforming. A remarkable exception is the
recycling processes. thermal cracking of polystyrene which yields approxi-
Melting of mechanically granulated plastic wastes in mately 70-90% of styrene monomer.9 The catalytic
order to reuse them in molding applications and energy cracking by contacting directly the polyolefins with a
recovered by incineration are at present the most catalyst, usually an acid solid, is an interesting alterna-
established alternatives for plastic recycling.1,2 How- tive, which allows to increase the conversion and to
ever, the former can be only applied to pure and high- obtain high-quality products. Different works have been
quality wastes whereas incineration suffers from a published using mainly amorphous SiO2-Al2O3 and
strong social rejection due to possible contribution of the different types of zeolites as catalysts.10-15 However,
effluent gases to the air pollution. due to the bulky nature of the plastic molecules, the
Chemical recycling of plastic wastes includes a num- cracking reaction is strongly controlled by the pore size
ber of processes and technologies directed toward their of the catalysts. Amorphous SiO2-Al2O3 usually present
conversion into the raw monomers or valuable feed- a wide distribution of pore radius whereas zeolites are
stocks, which are viewed as very promising alterna- microporous materials with pores in the range 0.4-0.8
tives.3 Most of the petrochemical companies are in-
volved in a variety of research projects for the develop-
(7) Kastner, H.; Kaminsky, W. Hydrocarbon Process. 1995, May,
ment of chemical routes of plastic recycling.4-6 For 109.
condensation polymers, such as polyesters, polyamides, (8) Wampler, T. P. J. Anal. Appl. Pyrol. 1989, 15, 187.
and polyurethanes, the starting monomers can be (9) Audisio, G.; Bertini, F. J. Anal. Pyrol. 1992, 24, 61.
(10) Uemichi, Y.; Kashiwaya, Y.; Tsukidate, M.; Ayame, A.; Kanoh,
H. Bull. Chem. Soc. Jpn. 1983, 56, 2768.
* To whom correspondence should be addressed. (11) Vasile, C.; Onu, P.; Barboiu, V.; Sabliovschi, M.; Moroi, G. Acta
X Abstract published in Advance ACS Abstracts, October 1, 1997. Polym. 1985, 36 (10), 543.
(1) Babinchak, S. CHEMTECH 1991, 21 (12), 728. (12) Vasile, C.; Onu, P.; Barboiu, V.; Sabliovschi, M.; Moroi, G.;
(2) Rowatt, R. J. CHEMTECH 1993, 23 (1), 56. Ganju, D.; Florea, M. Acta Polym. 1988, 39 (6), 306.
(3) Layman, P. Chem. Eng. News 1993, 71 (41), 11. (13) Ishihara, Y.; Nanbu, H.; Ikemaru, T.; Takesue, T. Fuel 1990,
(4) Hirota, T.; Fagan, F. N. Makromol. Chem., Macromol. Symp. 69, 978.
1992, 57, 161. (14) Audisio, G.; Bertini, F.; Beltrame, P. L.; Carniti, P. Makromol.
(5) Miller, A. Environ. Sci. Technol. 1994, 28, 16A. Chem., Macromol. Symp. 1992, 57, 191.
(6) Layman, P. Chem. Eng. News 1994, 72 (13), 19. (15) Lin, R.; White, R. L. J. Appl. Polym. Sci. 1995, 58, 1151.

S0887-0624(97)00055-8 CCC: $14.00 © 1997 American Chemical Society


1226 Energy & Fuels, Vol. 11, No. 6, 1997 Aguado et al.

nm, which hinders the access of the polymer molecules was dried at 110 °C overnight and activated by calcination at
into the zeolite channels and cavities. In fact, several 550 °C for 14 h.
authors have suggested that the initial steps of the Catalyst Characterization. The chemical composition of
polyolefin degradation over zeolites take place mainly the samples was determined by X-ray fluorescence (XRF) with
on the acid sites located on the external surface of the a Philips PW 1404 spectrometer. X-ray diffraction patterns
were collected with a Philips XPÄ ERT MPD diffractometer with
zeolite crystals.10,16 The same mechanism seems to
Cu KR radiation and Ni filter.
account for the reforming of heavy oil from waste High-resolution 27Al magic angle spinning nuclear magnetic
plastics over zeolites.17 resonance (MAS-NMR) spectra of the MCM-41 samples were
In this paper, following a previous work,18 we report recorded at 104.26 MHz, using a Bruker MSL-400 spectrom-
the results obtained in the catalytic cracking of low- and eter equipped with a Fourier transform unit. The spinning
high-density polyethylene (LDPE and HDPE, respec- frequency was 4000 cps with time intervals of 5 s between
tively) and polypropylene (PP) using MCM-41 as a successive accumulations. The measurements were carried
catalyst to yield liquid fuels. This material is a meso- out at room temperature with [Al(H2O)6]3+ as external stan-
porous silicate/aluminosilicate and first prepared in dard reference, the accumulations being amounted to 2000 and
1992 in the presence of alkyltrimethylammonium sur- 400 FIDs, respectively.
The N2 adsorption-desorption isotherms were measured at
factants.19,20 It is characterized by having uniform
77 K on a Micromeritics ASAP 2010 instrument using stan-
mesopores between 2 and 10 nm, whose size can be dard adsorption techniques. The samples were previously
adjusted by changing the synthesis conditions. There- outgassed by treatment at 200 °C for 5 h under vacuum. The
fore, MCM-41 is a potentially interesting catalyst for surface area was calculated using the BET equation whereas
the conversion of bulky substrates. The catalytic activ- the pore size distribution was determined by applying the BJH
ity and product distribution obtained with this material model with the Harkins and Jura equation assuming a
in the conversion of polyolefinic plastic into feedstocks cylindrical pore geometry. For ZSM-5 sample, the micropore
are discussed in this work by comparison to those volume and the external surface area were determined by
corresponding to amorphous SiO2-Al2O3 and ZSM-5 means of the t-plot method.
zeolite samples. NH3 thermal programmed desorption (TPD) measurements
were performed on a Micromeritics 2900 apparatus equipped
with a thermal conductivity detector. The samples were first
Experimental Section treated in an He stream at 560 °C and thereafter saturated
with NH3 at 180 °C for 30 min. The physically adsorbed NH3
Catalyst Preparation. The MCM-41 sample was synthe-
was removed by flowing He (50 mL(STP)/min) through the
sized according to a previously published procedure.20 Thereby,
sample at 180 °C for 90 min. Finally, the ammonia TPD was
a solution with 4.42 g of Cab-O-Sil silica and 1.67 g of sodium
carried out by increasing the temperature up to 550 °C with
hydroxide in 44 g of deionized water was added with stirring
a heating rate of 15 °C/min, the NH3 concentration in the
to a second one formed by 0.19 g of sodium aluminate, 7.48 g
effluent being continuously monitorized.
of cetyltrimethylammonium bromide (CTMABr), and 22 g of
deionized water. The mixture so obtained was loaded into a Catalytic Tests. The polyolefins used in this work were
Teflon-lined autoclave and kept at 120 °C for 2 days in static provided by REPSOL, and had the following features: LDPE
conditions. The synthesis product was separated by filtration, (average molecular weight, Mw ) 416 000), HDPE (Mw )
washed with deionized water, dried at 110 °C, and calcined 188 000), and PP (Mw ) 450 000, isotacticity index ) 93%).
at 550 °C in static air for 14 h. The acidic form of MCM-41 The thermal and catalytic cracking of these polyolefins were
was obtained by three times repeated ion exchange with 1 M performed at 400 °C and atmospheric pressure in a batch
NH4Cl aqueous solution at room temperature for 1 h, followed reactor with continuous N2 flow (25 mL(STP)/min). In each
by calcination at 550 °C. experiment, 1.2 g of plastic was loaded into the reactor and
mixed with the appropriate amount of catalyst. Previously,
ZSM-5 zeolite was prepared from ethanol-containing gels
the catalysts were ground and sieved until obtained a particle
at 170 °C for 24 h according to the method described else-
size below 74 µm. The reactor was heated to the desired
where.21 After the synthesis, the zeolite sample was ion-
reaction temperature in 15 min, which was kept constant for
exchanged with a 0.6 M HCl aqueous solution and then
a period of 30 min. During this time, the liquid and gaseous
calcined at 550 °C for 14 h.
products coming out from the reactor were separated in a
The amorphous silica-alumina was prepared by the sol-
condenser and accumulated to determine their composition by
gel route following a two-step method. In the first one, 16 g
gas chromatography. The gaseous products were analyzed
of tetraethyl orthosilicate was hydrolyzed with 10 g of 0.2 M
with a Hewlett-Packard 5880 GC on a Porapak Q column,
aqueous HCl at room temperature for 45 min. Once the
whereas the composition of the liquid products was determined
initially two phase system became monophasic, a solution
with a Perkin-Elmer 8310 GC using a 25 m long BP-5 capillary
containing 0.523 g of aluminum isopropoxide in 7 g of isopropyl
column.
alcohol was added and the mixture was stirred for 10 min to
Molecular Simulation. Molecular simulation techniques
complete the hydrolysis of the Si and Al alkoxides. In the
were applied to estimate the effective cross diameter of the
second step, the gel point was reached by dropwise addition
polymeric chains and to determine whether oligomeric frag-
of 21 wt % aqueous ammonia solution. The cogel so obtained
ments of the polyolefins used in this work can be adsorbed
and accommodated within the pore network of the ZSM-5
(16) Mordi, R. C.; Fields, R.; Dwyer, J. J. Chem. Soc., Chem.
Commun. 1992, 374. zeolite. The calculations were performed using the Cerius2
(17) Songip, A. R.; Masuda, T.; Kuwahara, H.; Hashimoto, K. Appl. program, developed by BIOSYM/Molecular Simulations.
Catal. B: Environ. 1993, 2, 153. The structure of oligomers with a chain length of 20 carbon
(18) Aguado, J.; Serrano, D. P.; Romero, M. D.; Escola, J. M. Chem. atoms was generated with a molecular mechanics force field,
Commun. 1996, 725.
(19) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, universal force field (UFF), including a parametrized descrip-
J. S. Nature 1992, 359, 710. tion of the full periodic table, which has been proved to
(20) Beck, J. S.; Vartuli, J. C.; Roth, W. J.; Leonowicz, M. E.; Kresge, reproduce the structure of many organic and inorganic com-
C. T.; Schmitt, K. D.; Chu, C. T.-W.; Olson, D. H.; Sheppard, E. W.; pounds.22,23 The partial charges in the atoms of the molecules
McCullen, S. B.; Higgins, J. B.; Schlenker, J. L. J. Am. Chem. Soc.
1992, 114, 10834.
(21) Uguina, M. A.; de Lucas, A.; Ruiz, F.; Serrano, D. P. Ind. Eng. (22) Rappé, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard III, W. A.;
Chem. Res. 1995, 34, 451. Skiff, W. M. J. Am. Chem. Soc. 1992, 114, 10024.
Conversion of Polyolefins into Liquid Fuels over MCM-41 Energy & Fuels, Vol. 11, No. 6, 1997 1227

were calculated by a charge equilibration method,24 followed


by an energy minimization process in order to determine the
most stable configuration of the molecule. The effective
diameters of the cross section in the oligomers were estimated
as the distance between the atomic centers plus their van der
Waals radii. This diameter was calculated at the points in
the oligomeric chain having the maximum cross section.
Molecular simulations oligomers adsorption on the ZSM-5
structure were performed at fixed pressure (101.33 kPa),
temperature (673 K), and volume in the grand canonical
ensemble by the Monte Carlo technique (GCMC).25 A number
of 12-unit cells of MFI topology were taken as simulation box.
The interactions adsorbate-adsorbent and adsorbate-adsor-
bate were modeled by Lennard-Jones 12-6 potential and the
values of the corresponding parameters were taken from the
literature.26 Partial charges were also placed in the zeolite
atoms (Si +2; O -1) and Coulombic interactions were also
taken into account. The average loading of the oligomers on
the ZSM-5 unit considered was determined with 107 configura-
tions.

Results and Discussion


Physicochemical Properties of the Catalysts.
The different catalysts used in this work for the
degradation of polyolefinic plastics were characterized
by several techniques in order to gain information about
the properties which are mainly responsible for their
catalytic activity: porous structure, state of the Al
atoms, and concentration and strength of the acid sites.
This characterization was carried out in more detail on
the MCM-41 sample, since this is a new and less studied
material.
XRD spectra of the as-synthesized MCM-41 catalyst
exhibit a single strong reflection corresponding to the
d100 spacing at 3.9 nm, which is typical of this kind of
mesoporous materials.19 TG analysis in air of the as-
synthesized MCM-41 leads to a 49% weight loss between
Figure 1. N2 adsorption at 77 K on the MCM-41 and the
200 and 550 °C due to the decomposition and removal amorphous SiO2-Al2O3 samples: (a, top) adsorption isotherms,
of the surfactant molecules occluded within the pores. (b, bottom) pore size distribution.
This value is in agreement with the data previously
reported in the literature.27 The maximum of the major Table 1. Properties of the Catalysts
peak in the differential curve is located at ca. 250 °C, N2 adsorption (77 K) NH3 TPD (180-550 °C)
showing that the surfactant can be completely removed SBET VP acidity Tmax
by calcination of the sample at 550 °C. After this catalyst Si/Al (m2/g) (cm3/g)a (mmol/g) (°C) NH3/Al
thermal treatment, the position of the XRD reflection MCM-41 42.7 1327 1.51 0.30 314 0.79
decreases by 0.35 nm due to a contraction of the SiO2-Al2O3 35.6 261 0.97 0.24 302 0.53
structure. ZSM-5 31.0 362 0.18 0.52 470 1.00
The N2 adsorption isotherm at 77 K and the pore size a Total pore volume measured at p/p0 ) 0.995.
distribution of the MCM-41 are compared in Figure 1,
a and b, respectively, to those of the amorphous SiO2- area of the amorphous SiO2-Al2O3 sample is ap-
Al2O3. The surface area and pore volume of the cata- preciably lower, exhibiting a very wide distribution of
lysts are given in Table 1. For the MCM-41 sample, pore sizes. In fact, this material presents pores with
most of the N2 adsorption takes place at relative sizes in the whole range between 4 and 110 nm, with
pressures ranging between 0.2 and 0.6, with an inflec- three maxima around 10, 45, and 100 nm, denoting the
tion point at p/p0 ) 0.3 which is a typical feature of the irregularity of its pore structure.
uniform mesoporosity of this material.19 The high The Al content of the catalysts is also given in Table
surface area of MCM41 is remarkable, while the pore 1. The Si/Al ratio of the MCM-41 sample is higher than
size distribution shows the presence of well-defined the one corresponding to the starting synthesis gel
pores with 2.9 nm diameter. In contrast, the surface (Si/Al ) 30), showing the incorporation of the Al atoms
into this structure is less favored than that of the Si
(23) Casewit, C. J.; Colwell, K. S.; Rappé, A. K. J. Am. Chem. Soc. species. An important matter is the state of the Al
1992, 114, 10035. atoms in the MCM-41 sample since different authors
(24) Rappé, A. K.; Goddard III, W. A. J. Phys. Chem. 1991, 95, 3358.
(25) Allen, M. P.; Tildesley, D. J. Computer Simulations of Liquids;
have reported that, depending on the Al source used in
Oxford University Press Inc.: New York, 1996. the synthesis and/or the activation conditions, octa-
(26) Yashonath, S.; Thomas, J. M.; Novak, A. K.; Cheetham, A. K. hedral extraframework Al species can be formed.27-29
Nature 1988, 311, 601.
(27) Chen, C. Y.; Li, H. X.; Davis, M. E. Microporous Mater. 1993, Figure 2 illustrates the 27Al MAS NMR spectra of both
2, 17. the as-synthesized and the activated MCM-41 samples.
1228 Energy & Fuels, Vol. 11, No. 6, 1997 Aguado et al.

curves corresponding to the MCM-41 material and the


amorphous SiO2-Al2O3 sample with desorption maxi-
mum at 314 and 302 °C, respectively, which are quite
lower than the one observed in the ZSM-5 zeolite (470
°C). These results confirm that the acid sites in MCM-
41 are weaker than those of ZSM-5 and very close in
strength to the acid sites of amorphous materials. In
fact, it has been reported that, in spite of this ordered
and regular porosity, the pore walls of MCM-41 are
amorphous, without any crystalline order of the SiO4
and AlO4- tetrahedra.27
The ratio between the adsorbed-desorbed NH3 and
the Al content varies greatly among these samples. For
ZSM-5 this ratio is unity, indicating that all the Al
atoms are associated to acid sites accessible to the
ammonia molecules. The values of this ratio are 0.79
and 0.53 for MCM-41 and the amorphous SiO2-Al2O3
sample, respectively. These differences may be due to
several reasons concerning the MCM-41 acidity: (a)
Figure 2. 27Al MAS NMR spectra of the MCM-41 before and presence of extraframework Al species, (b) existence of
after activation.
very weak acid sites that release the adsorbed NH3
below the starting temperature in the TPD experiment
(180 °C), (c) different nature of the acid sites present in
each sample, and (d) location of a part of the Al atoms
within the pore walls, in positions inaccessible to the
NH3 molecules. In this way, it has been recently
reported that the acid sites present in MCM-41 materi-
als are both Brönsted and Lewis sites with a high
proportion of the latter,30-32 whereas it is known that
the acidity of ZSM-5 samples is mainly related to
Brönsted sites. The adsorption of ammonia on these
two types of acid sites may proceed with a different
stoichiometry and/or take place in a different temper-
ature range, which could explain the observed variation
in the ratio between the adsorbed NH3 and the Al
content.
The width of the pore wall in MCM-41 can be
estimated from both XRD and N2 adsorption data.
According to the XRD d100 spacing, the hexagonal unit
cell parameter (a0 ) 2d100/x3) of the calcined MCM-41
Figure 3. NH3 TPD of the ZSM-5, MCM-41, and amorphous sample is 4.1 nm which, after subtraction of the pore
SiO2-Al2O3 samples. diameter, leads to a thickness of the pore walls of 1.2
nm. This result agrees well with values earlier reported
In the first case, an unique peak with a chemical shift in the literature for MCM-41 materials. Thus, Chen et
of 53 ppm is observed proving that all the Al atoms are al.33 have concluded that a thickness of 1.0 nm is enough
tetrahedrally coordinated in the as-synthesized sample. to include between two and three monolayers of TO4
However, after the activation of the sample by calcina- tetrahedra along the pore wall. Likewise, a model for
tion at 550 °C to remove the surfactant, ion exchange MCM-41 has been developed by molecular dynamics
of Na+ by NH4+, and final calcination at 550 °C to simulation34 with wall thickness around 1 nm, showing
generate the acid form, a peak at 0 ppm appears, which the presence of several TO4 tetrahedra along the wall,
denotes the presence of Al atoms with octahedral the middle ones being not in contact with the pores. If
coordination, probably due to the formation of ex- it is assumed that the Al atoms are randomly distrib-
traframework species. Nevertheless, the proportion of uted along the pore wall, it means that a percentage of
octahedral Al in the calcined protonic MCM-41 is them will not be accessible to the NH3 molecules. We
estimated to be just around 5% from the corresponding think this is the major reason for the discrepancy
peak area in the NMR spectra. observed between the NH3 TPD measurements and the
The acidic properties of the catalysts were tested by Al content of MCM-41, in addition to the presence of
NH3 TPD between 180 and 550 °C. The desorption around 5% of extraframework Al. In the case of the
curves obtained are compared in Figure 3, while the
amount of adsorbed-desorbed ammonia and the position (30) Busio, M.; Jänchen, J.; van Hooff, J. H. C. Microporous Mater.
of the TPD maximum are shown in Table 1. It is 1995, 5, 211.
(31) Corma, A.; Grande, M. S.; Gonzalez-Alfaro, V.; Orchilles, A. V.
interesting to note the similarity between the TPD J. Catal. 1996, 159, 375.
(32) Mokaya, R.; Jones, W.; Luan, Z.; Alba, M. D.; Klinowski, J.
(28) Janicke, M.; Kumar, D.; Stucky, G. D.; Chmelka, B. F. Stud. Catal. Lett. 1996, 37, 113.
Surf. Sci. Catal. 1994, 84, 243. (33) Chen, C. Y.; Burkett, S. L.; Li, H. X.; Davis, M. E. Microporous
(29) Luan, Z.; Cheng, C. F.; Zhou, W.; Klinowski, J. J. Phys. Chem. Mater. 1993, 2, 27.
1995, 99, 1018. (34) Feuston, B. P.; Higgins, J. B. J. Phys. Chem. 1994, 98, 4459.
Conversion of Polyolefins into Liquid Fuels over MCM-41 Energy & Fuels, Vol. 11, No. 6, 1997 1229

Figure 4. Carbon distribution in the products of the HDPE Figure 5. Carbon distribution in the products of the LDPE
cracking (400 °C, 0.5 h, plastic/catalyst ) 18 w/w). cracking (400 °C, 0.5 h, plastic/catalyst ) 18 w/w).
Table 2. Catalytic Cracking of HDPE (400 °C, 0.5 h, contrary, the mesopores present in MCM-41 and the
plastic/catalyst ) 18 w/w) amorphous SiO2-Al2O3 should favor the diffusion and
product selectivity (%) subsequent reaction of polyethylene. However, the
conv C1-C4 C2-C4 order of activities observed suggests that the major
catalyst (%) paraf olef C5-C12 C13-C22 C23-C40 factor influencing the polyethylene cracking is the
MCM-41 35.2 14.6 20.0 52.6 12.6 0.2 acidity rather than the pore size of the catalysts. The
SiO2-Al2O3 9.2 15.9 19.6 50.1 13.1 1.3 most active catalyst is the ZSM-5 zeolite, which presents
ZSM-5 96.4 13.2 38.9 46.3 1.5 0.1 the highest and strongest acidity. On the other hand,
the superior activity of MCM-41 compared to the
Table 3. Catalytic Cracking of LDPE (400 °C, 0.5 h,
plastic/catalyst ) 18 w/w) amorphous SiO2-Al2O3 can be mainly assigned to the
higher surface area of the former, since both materials
product selectivity (%)
present similar acid sites.
conv C1-C4 C2-C4 It is interesting to note that the differences in
catalyst (%) paraf olef C5-C12 C13-C22 C23-C40
catalytic activity among the three samples are shortened
MCM-41 67.6 6.9 12.6 63.9 16.3 0.3 when going from HDPE to LDPE. For the same
SiO2-Al2O3 34.4 9.4 14.9 55.8 18.9 1.0 reactions conditions, the catalytic cracking of LDPE
ZSM-5 95.4 19.8 32.8 44.6 2.3 0.5
proceeds faster than the degradation of HDPE. This
amorphous SiO2-Al2O3, its lower surface area must lead effect is more evident when using MCM-41 and the
to a lower proportion of accessible Al atoms, explaining amorphous SiO2-Al2O3 as catalysts and it is a conse-
the decrease in the NH3/Al ratio compared to MCM-41. quence of the structure of these polymers. High-density
Catalytic Cracking of Polyethylene. The catalytic polyethylene is typically formed by linear macromol-
degradation of both high- and low-density polyethylene ecules, whereas low-density polyethylene is character-
has been investigated at 400 °C in a batch reactor with ized by a certain degree of branching. The presence of
a duration of the experiments of 30 min. Under these tertiary carbons in LDPE provides favorable positions
conditions, the conversions obtained in blank experi- for the initiation of the polymer chain cracking since
ments without catalyst were almost negligible (less than their activation, by hydride abstraction or by addition
1%), showing that the thermal cracking of these poly- of a proton, does not require a very strong acidity.
mers at 400 °C is very slow. In both thermal and These reactions can be easily catalyzed by mild acid
catalytic tests, the plastic conversion values have been sites, such as those present in MCM-41 and amorphous
calculated by considering only those products having SiO2-Al2O3 samples.
boiling points low enough to leave the reactor in the N2 Great differences are also observed among the cata-
stream, which accounts for hydrocarbons with a total lysts regarding the product distribution. Over ZSM-5
number of carbons up to approximately 40. Heavier zeolite the cracking of both HDPE and LDPE leads to a
cracking products, remaining in the reactor, were not high proportion of gaseous hydrocarbons rich in olefins.
considered in calculating the plastic conversion. In contrast, the main products obtained with MCM-41
Tables 2 and 3 summarize the results obtained in the and amorphous SiO2-Al2O3 are liquid fractions with
catalytic cracking of HDPE and LDPE over the different boiling points in the range of gasolines (C5-C12) and
catalysts, while the product distributions per carbon middle distillates (C13-C22), suggesting that the product
atom number are shown in Figures 4 and 5, respec- selectivities are well correlated with the pore size
tively. For both polymers, the highest activity is distribution of each catalyst. In this way, it must be
observed with the ZSM-5 zeolite with conversions close pointed out that the upper limit observed for the product
to 100%. This catalyst has pores with diameters around distribution corresponding to the polyethylene cracking
0.55 nm, and hence it was expected that the access and over ZSM-5 (C12) is close to the largest molecules which
intracrystalline diffusion of the bulky plastic molecules are typically formed within its pore system. This fact
to the internal acid sites are strongly hindered. On the suggests that over the ZSM-5 sample used in this work,
1230 Energy & Fuels, Vol. 11, No. 6, 1997 Aguado et al.

Table 4. Catalytic Cracking of PP (400 °C, 0.5 h, plastic/


catalyst ) 36 w/w)
product selectivity (%)
conv C1-C4 C2-C4
catalyst (%) paraf olef C5-C12 C13-C22 C23-C40
MCM-41 99.2 3.9 10.7 65.0 20.0 0.4
SiO2-Al2O3 47.9 4.0 9.8 51.5 34.1 0.6
ZSM-5 11.3 22.3 27.0 50.0 0.7 0.0

the polyethylene cracking takes place mainly within the


pores rather than on the external surface of the crystals.
Two maxima are present in the product distribution
per carbon atom number obtained with the three
catalysts investigated. In the HDPE catalytic conver-
sion, the first maximum is observed for C4 hydrocarbons
over the three catalysts, whereas in the degradation of
LDPE over MCM-41 and amorphous SiO2-Al2O3 it is a
less pronounced maximum and displaced to C5 species.
According to earlier works, the catalytic degradation of
Figure 6. Carbon distribution in the products of the PP
polyolefins proceeds through two main ionic mecha- cracking (400 °C, 0.5 h, plastic/catalyst ) 36 w/w).
nisms:11,15,35,36 (i) cracking at the end groups of the
polymeric chains, and (ii) random cleavage of the than those observed for HDPE and LDPE in spite of
polymer molecules at any bond in the chain. In both using a plastic/catalyst ratio 2-fold higher.
cases, the starting molecules may be intact polymer All these results are related and can be explained by
molecules or partially cracked polyolefins (oligomers). the existence of a high proportion of tertiary carbons in
Simultaneously to the catalytic degradation, the ther- the polypropylene chains, which are very reactive to
mal cracking of the plastic molecules into oligomers may undergo both thermal and catalytic cracking reactions.
also take place through a radical mechanism. C3-C5 In fact, half of the carbon atoms in the polypropylene
hydrocarbons are the main products expected from the chains are tertiary due to the presence of side-chain
first catalytic pathway, which originates the first maxi- methyl groups.
mum present in the carbon distribution. The rest of the An important conclusion can be derived from the
products observed can be formed by random cleavage practical absence of activity of the ZSM-5 sample
of any bond in the chain or from the initial products of compared to just the thermal cracking, since it can be
the end chain cracking through secondary reactions assigned to steric and/or diffusion hindrances. The
such as oligomerization, aromatization, etc. Thus, since presence of the side-chain methyl groups increases the
the second maximum of the carbon distribution is effective cross section of the polypropylene molecules
observed in the range C6-C9, it is probably originated compared to the polyethylenic chains, which may pre-
by the oligomerization of the C3-C5 olefins. vent their access to the active sites located within the
zeolite pores. By molecular simulation on oligomeric
Catalytic Cracking of Polypropylene. The cata-
units formed by 20 chain carbons, it has been possible
lytic conversion of polypropylene has been also inves-
to calculate the effective cross-section diameter of these
tigated at 400 °C, as in the case of polyethylene. One
molecules. We have taken as effective diameter the
significant difference between the thermal behaviour of
maximum atom distance measured perpendicularly to
these two plastics is that the thermal degradation of
the longitudinal chain axis. The results obtained are
polypropylene is faster, as denoted by a certain conver-
as follows: 0.55 nm for the polyethylene oligomer, 0.64
sion (around 8%) observed in experiments without
nm for the isotactic polypropylene oligomer, and 0.73
catalyst after 30 min of reaction.
nm for the syndiotactic polypropylene oligomer. Com-
The results obtained in the catalytic degradation of pared to the crystallographic pore size of the ZSM-5
polypropylene are shown in Table 4 and Figure 6. In zeolite (0.55 nm), those numbers show the difficulty for
this case the order of activity is clearly different from both types of polypropylene to enter and diffuse through
that corresponding to polyethylenic plastics. Thus, the the zeolite channels. Nevertheless, it has to be taken
MCM-41 sample leads to almost 100% conversion into account that neither the molecules nor the zeolite
whereas the activity obtained with the ZSM-5 zeolite structure are completely rigid but they can be deformed
is very close to that of the thermal cracking. Even the to certain extension by thermal vibration, which ex-
conversion with the amorphous SiO2-Al2O3 is quite plains that molecules somewhat larger than the crystal-
higher than the observed with the zeolite in spite of the lographic pore diameter may have access to the internal
highest surface area and stronger acidity of the latter. zeolite volume. In order to address this possibility, the
Moreover, as has been commented for the thermal adsorption of the polymer oligomers on the MFI struc-
cracking, the polypropylene conversion over MCM-41 ture at different temperatures has been predicted by
and amorphous SiO2-Al2O3 is quite faster than the molecular simulation. The results obtained show that
degradation of both high- and low-density polyethylene. the C20 polyethylene oligomer can be accommodated
Note that these catalysts present PP conversions higher within the straight ZSM-5 channels, whereas the ad-
sorption capacity for both polypropylene oligomers is
(35) Ivanova, S. R.; Gumerova, E. F.; Minsker, K. S.; Zaikov, G. E.; zero, confirming the latter cannot enter the zeolite pores.
Berlin, A. A. Prog. Polym. Sci. 1990, 15, 193.
(36) Ishihara, Y.; Nanbu, H.; Saido, K.; Ikemura, T.; Takesue, T. At the same time, this result implies that the contri-
Bull. Chem. Soc. Jpn. 1991, 64, 3585. bution to the plastic cracking of the acid sites present
Conversion of Polyolefins into Liquid Fuels over MCM-41 Energy & Fuels, Vol. 11, No. 6, 1997 1231

on the external surface of the crystals in the ZSM-5 to the amorphous SiO2-Al2O3, which is interpreted in
sample used in this work is negligible. Therefore, the terms of its higher surface area and more regular
high conversion obtained with this material in the mesoporosity. The differences in conversion among the
HDPE and LDPE cracking takes place by penetration catalysts are reduced when going from HDPE to LDPE
of polymer or oligomer molecules, at least up to a certain due to the presence of a higher proportion of tertiary
chain length, into the zeolite pores and subsequent carbons in the second polymer, which favors the initia-
reaction over the internal acid sites. This conclusion tion of the chain cracking. On the other hand, the
agrees with the above commented relationship between product distribution obtained with each catalyst is
the largest compounds observed in the product distribu- determined mainly by its pore size distribution. Over
tion of HDPE and LDPE cracking over ZSM-5 and the ZSM-5, almost 50% of the products from both HDPE
maximum size of the molecules that are usually formed and LDPE cracking are gaseous hydrocarbons (C2-C4)
within the zeolite pores. On the contrary, if the initial with a high proportion of olefins, whereas in the case
cracking takes place on the zeolite external acid sites, of MCM-41 and amorphous SiO2-Al2O3 the degradation
a wider product distribution should be observed and the of these polymers leads mainly to liquid fractions with
polypropylene cracking should proceed in a higher boiling points in the range of gasolines (C5-C12) and
extension. Nevertheless, it has to be taken into account middle distillates (C13-C22).
that the relative proportion and the catalytic contribu- In the polypropylene degradation, the order of activity
tion of the external surface of zeolites greatly depends is strongly modified compared to the polyethylenic
on the zeolite crystal size. In our case, the ZSM-5 plastics. The highest activity is observed with the
sample presents an average crystal size around 5 µm MCM-41 sample, while the conversion obtained with the
and, according to the t-plot method applied to the N2 ZSM-5 zeolite is very close to that of the thermal
adsorption isotherm, it has 7.1 m2/g of external surface cracking. With MCM-41 and the amorphous SiO2-
area, which means that approximately just a 2% of the Al2O3 the conversion of polypropylene is faster than in
total surface area is external. It cannot be discarded the case of HDPE and LDPE, which is related to the
that, if ZSM-5 samples with lower crystal sizes are used, existence of a high proportion of tertiary carbons in the
the contribution of the external acidity to the polyolefins former. The product distribution of the polypropylene
cracking becomes significant. cracking over these catalysts is qualitatively similar to
Concerning the product distribution of the poly- those of HDPE and LDPE with liquid hydrocarbon
propylene cracking, the results show similar trends to mixtures as predominant products.
those obtained in the polyethylene conversion. MCM-
The plastic cracking over the ZSM-5 sample used in
41 and amorphous SiO2-Al2O3 lead mainly to liquid
this work proceeds mainly within the zeolite pores
mixtures of hydrocarbons in the range of gasolines and
rather than on the external surface, as denoted by the
middle distillates, with a higher proportion of C13-C22
product distribution with hydrocarbons up to C12 ob-
products with the second catalyst, whereas C3-C5
tained in the conversion of HDPE and LDPE and the
hydrocarbons are the main products formed in the
absence of activity compared to the thermal cracking
experiment carried out with the ZSM-5 zeolite. Like-
in the degradation of PP. For both isotactic and
wise, two maxima at C5 and around C8-C9 are present
syndiotactic polypropylene, molecular simulation mea-
in the carbon distribution corresponding to both MCM-
surements indicate that the increase in their cross
41 and amorphous SiO2-Al2O3 samples. In the last
sections compared to polyethylene, derived from the
case, the first maximum is less pronounced while the
presence of methyl side groups, hinders their access to
distribution in the range C8-C18 is more uniform, which
the zeolite microporosity. Nevertheless, it cannot be
suggests a lower contribution for this catalyst of the
discarded that, if ZSM-5 samples with small crystallites
cracking reactions taking place at the end of the chains.
are used, the contribution of the external acid sites
It can be interpreted as a consequence of the presence
becomes significant.
in the amorphous SiO2-Al2O3 of larger pores with
irregular sizes and shapes which, besides the existence All these results show that MCM-41 is a material with
of a high proportion of highly reactive tertiary carbons, promising catalytic properties for the conversion of
favors the random cracking at any position in the chain. polymeric wastes into liquid feedstocks. It presents a
On the contrary, the MCM-41 sample is formed by one- high surface area and, although its acidity is weaker
dimensional and regular pores with diameters around than that of zeolites, the access of the polymer molecules
2.8 nm. Then, the plastic and/or oligomer molecules are to the active sites is not hindered as in the latter.
probably accommodated in a linear configuration along
the pores, exposing preferably the end of the chains to Acknowledgment. This work has been funded by
the active sites as they diffuse through the channels. the Comisión Interministerial de Ciencia y Tecnologı́a
from Spain (Projects CICYT AMB-94/0681 and MAT-
Conclusions 95/2044E). The authors are also grateful to REPSOL
S.A. for providing the plastic samples and to the
In the catalytic cracking of HDPE and LDPE, the
Research Center of CEPSA in Madrid for the XRF
highest activity is observed with the ZSM-5 zeolite due
measurements.
to its stronger acidity. The conversions exhibited by the
MCM-41 sample are superior than those corresponding EF970055V

You might also like