Вы находитесь на странице: 1из 10

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Environmental Sciences 20 (2014) 215 – 224

4th International Conference on Sustainable Future for Human Security, SustaiN 2013

Fuel Production from LDPE Plastic Waste over Natural Zeolite


Supported Ni, Ni-Mo, Co and Co-Mo Metals
Wiwin Sriningsih, Monica Garby Saerodji, Wega Trisunaryanti*, Triyono,
Ria Armunanto, Iip Izul Falah
Department of Chemistry, Universitas Gadjah Mada, Bulaksumur, Yogyakarta 55281, Indonesia

Abstract

Hydrocracking of LDPE plastics into fuel over bi-functional catalysts was systematically studied. The natural zeolite from
Sukabumi, Indonesia was refluxed in HCl 6N for 30 minutes and dried in an oven, then the activated natural zeolite (Z) was
produced. This study found that natural zeolite obtained from Sukabumi which consists of the mordenite type crystalline. For bi-
functional catalysts, Ni, Ni-Mo, Co and Co-Mo metals were loaded into the activated natural zeolite respectively to increase the
activity and selectivity for hydrocracking process. The catalyst samples were heated in a microwave and done the reduction by
hydrogen gas stream, produced Ni/Z, Ni-Mo/Z, Co/Z and Co-Mo/Z. Hydrocracking process was carried out effectively at 350 °C
with 20 mL/minutes of hydrogen gas stream for 1 hour of reaction time. Conversion of LDPE plastic waste over Co-Mo/Z at
350 °C produced the higher selectivity of gasoline, reached 71.49%. Based on the GCMS data, liquid yield consisted of
hydrocarbon compounds with atom chain of C6 until C19 that indicate fuel chemical structures, such as paraffins, olefins, and
naphthenes. By using this technology, the conversion of plastic waste into fuel is expected to reduce the environmental pollution,
support the use of soil, and increase the energy storage.

© 2013
2014 The
WegaAuthors. Published
Trisunaryanti. by Elsevier
Published B.V. B.V. Open access under CC BY-NC-ND license.
by Elsevier
Selection and peer-review under responsibility
responsibility ofthe
of theSustaiN
SustaiNconference
conferencecommittee
committeeand
andsupported
supportedby
byKyoto
KyotoUniversity;
University;(RISH),
(RISH),
(OPIR), (GCOE-ARS)
(OPIR), (GCOE-ARS) and and (GSS)
(GSS) as
as co-hosts.
co-hosts

Keywords: LDPE plastic; natural zeolite; metals; hydrocracking; fuel

* Corresponding author. Tel.: +628-112-829-85; fax: +62-274-545-188.


E-mail address:wegats@ugm.ac.id

1878-0296 © 2014 Wega Trisunaryanti. Published by Elsevier B.V. Open access under CC BY-NC-ND license.
Selection and peer-review under responsibility of the SustaiN conference committee and supported by Kyoto University; (RISH), (OPIR),
(GCOE-ARS) and (GSS) as co-hosts
doi:10.1016/j.proenv.2014.03.028
216 Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224

1. Introduction

Plastic wastes, both from industries and households have increased quite bombastic. Worldwide plastics
production rose to 280 million tons in 2011, according to first rough estimates published by Plastics Europe. From
2010 to 2016, global plastics consumption is expected to grow by an average of about 4% each year. It will be
equivalent to the amount of waste generated. The main components of the household plastic waste streams include
the following five families of plastics: polyethylene (low-density polyethylene (LDPE), linear low-density
polyethylene (LLDPE), high-density polyethylene (HDPE)), polypropylene (PP), polyvinyl chloride (PVC),
polystyrene (PS), and polyethylene-terephthalate (PET), which are accounted entirely for 74% of all plastic wastes
[1].
Plastics are non-biodegradable materials so these wastes constitute various complicated problems, because of the
loss of natural resources, the environmental pollution, and the depletion of landfill space. Plastic waste management
is intended to decrease the environmental impact. The current situation might be described as a search of mature
technologies that can eliminate and process these plastic wastes with the lowest environmental impact and the
highest possible profitability. There are many ways of plastic waste management, such as: reducing, reusing,
mechanical recycling, incineration, energy recovery, and landfilling. Reducing and reusing are options with limited
applicability. Landfilling and burning plastics in incinerator will cause pollution, because they produce toxic gas,
such dioxins and furans [15]. Hence, recycling and energy recovery are alternatives to be also considered [2]. The
popular energy recovery methods are gasification and catalytic thermal degradation [3–5].
Gasification of plastic wastes into synthesis gas has been proposed as one of these technologies, but it is very
costly and requires the construction of large plants to be profitable. Cracking technologies toward transportation
fuels (gasoline, diesel) and chemicals are more flexible and receive more attention, in a context of growing prices of
crude oil [11]. Thermal cracking demands relatively high temperatures and its products require further processing for
their quality to be upgraded, so the method is less desirable. On the other hand, catalytic cracking of plastic waste
offers considerable advantages. It operates at considerably lower temperatures, since it reduces the activation energy
of the cracking. The use of catalysts will be more beneficial because it can lead to lower operating conditions of the
cracking process. It requires lower operating costs as well [6]. Furthermore, the catalyst allows the selectivity of the
plastic conversion to be tailored so it can be targeted toward the desired products (e.g. gases, gasoline, or diesel) just
by choosing adequately the acidity and pore structure of the catalyst [11].
The hydrocracking process, which basically converts high-boiling molecules into more desirable lower molecular
weight products by simultaneous or sequential hydrogenation and carbon bond breaking, is not only an important
process used in modern oil refineries, but also one of the most promising processes for conversion of plastic wastes.
There have been a number of investigations in the recent past concerning the use of solid acid catalysts for the
degradation of plastic wastes into fuel oils. These catalysts were selected because of their excellent activities and
stabilities [4].
Bi-functional catalysts are characterized by the presence of both acid and metal sites. The acid sites are usually
provided by alumina-silicate support (zeolites, MCM-41, and amorphous silica-alumina), while the metal sites come
from a supported metal, such as Ni, Mo, W, Co, or their combinations, oxide, or sulfide of group VIII and/or group
VI B. The acidic supports used in today’s modern petroleum industry are mostly silica-alumina [3,19,20] and various
zeolites, such as ZSM-5 [4], Y [13,17,18], and beta [5]. However, the use of these synthetic catalysts takes the
expensive cost and requires the preparation techniques with the high precision and accuracy.
Natural zeolite is recently reported as excellent material to support metal in the preparation of catalyst. Natural
zeolite materials are easily found in USA, Japan, Cuba, Soviet Union, Italy, Czechoslovakia, Hungary, Bulgaria,
South Africa, Yugoslavia, Mexico, Korea, and Indonesia. In Indonesia, natural zeolite reserves were discovered
more than 205.82 million tons. Zeolite deposits are most numerous in the Java island,i.e: Central Java (Wonosari,
Klaten), West Java (Bogor, Tasikmalaya, Sukabumi), and East Java (Bayah). Hydrocracking of plastic waste has
been widely studied in literatures [1–6,11,12]. However, the Indonesia’s natural zeolite especially from Sukabumi is
rarely applied as a catalyst for hydrocracking of LDPE plastic waste. Trisunaryanti et al. have observed the natural
zeolite from Sukabumi. It contains the most mordenite mineral that has high thermal stability. Hence, this work
would study the effect of temperature condition as well. The activation through acid treatment and modification
using supported metals could increase its acidity, pore and surface area as well could produce the good catalyst for
Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224 217

hydrocracking processes. Furthermore, the natural zeolite is very abundant and cheap, hence its use as a catalyst is
valuable to lower the cost of production [6].
Hydrocracking of plastic waste into fuel shows the highest potential for a successful future commercial plastic
recycling process, especially as we can consider that plastic waste is a cheap source of raw materials in times of
accelerated depletion of natural resources [15,16].

2. Experimental section

2.1. Treatment of natural zeolite

Natural zeolite was obtained from Sukabumi, West Java province of Indonesia. All of reagents were pro analyzed
grade from E. Merck (Germany), i.e.: hydrochloric acid (HCl) 6N, ammonium chloride (NH 4Cl) 1N, ammonium
hydroxide (NH4OH) 25%, and pyridine. Distilled water was provided by Basic Chemistry Laboratory, Department
of Chemistry UGM. Hydrogen gas was obtained from Samator Gas. Natural zeolite was soaked in distilled water for
3 days; afterwards it was dried in an oven at 70 °C for 4 hours. Zeolite was crushed and sieved to 250 mesh sieve. A
hundred grams of natural zeolite and 200 mL of HCl 6N was refluxed for 30 minutes at 50 °C, filtered and washed
with distilled water repeatedly until none of Cl- ions was detected, then dried at 70 °C for 4 hours in an oven. These
samples were soaked in NH4Cl 1N for 1 week, stirred regularly, decanted, washed with distilled water, dried, and
brayed to 100 mesh size. This sample was referred as an activated natural zeolite (Z).

2.2. Metals impregnation of the catalytic support

The metal salts, nickel-(II) nitrate hexahydrate (Ni(NO3)2.6H2O) 99.95%, ammonium molybdatetetrahydrate
((NH4)6Mo7O24.4H2O) 99.98%, and cobalt-(II) nitrate hexahydrate (Co(NO3)2.6H2O) 99.95% were purchased from
E. Merck (Germany). For monometal catalysts, Z sample was refluxed with a metal salt solution of Ni(NO3)2.6H2O
at 90 °C for 4 hours. Amount of Ni metal supported was 1wt% of Z sample. The result was evaporated and dried in
an oven at 120 °C for 2 hours in order to obtain Ni/Z catalyst. Co/Z catalyst was made using Co(NO 3)2.6H2O
precursor with the same way as Ni/Z. For bimetal catalysts, Z sample was refluxed with a metal salt solution of
(NH4)6Mo7O24.4H2O at 90 °C for 4 hours. Amount of Mo metal supported was 0.5wt% of Z sample. The result was
evaporated and dried in an oven at 120 °C for 2 hours in order to obtain the Mo/Z catalyst. It was refluxed with a
metal salt solution of Ni(NO3)2.6H2O at 90 °C for 4 hours. Amount of Ni metal supported was 0.5wt% of Z sample.
The result was evaporated and dried in an oven at 120 °C for 2 hours in order to obtain the Ni-Mo/Z catalyst. Co-
Mo/Z catalyst was made with the same way as Ni-Mo/Z using Co(NO3)2.6H2O and (NH4)6Mo7O24.4H2O precursors.
Catalyst samples were heated in a microwave at 450 watt for 30 minutes and followed by reducing under the
hydrogen gas stream (20 mL/minutes) at 400 °C for 3 hours. Determination of the effect of modified treatments
towards dealumination, the crystallinity of catalyst samples and the total acid site amount on the catalyst samples
were analyzed by using FT-IR Spectroscopy Shimadzu Prestige-21, Shimadzu XRD-6000, and gravimetrically using
ammonia as well as pyridine base vapour adsorption, respectively.

2.3. Preparation of feed

Low-density polyethylene (LDPE) plastics were collected from the household waste. The collected plastics were
washed and dried under the sunshine for 12 hours. The dried plastics were vaporized in a reactor at 500 °C and
condensed into liquid fraction.

2.4. Hydrocracking processes

Hydrocracking processes were carried out under the hydrogen gas stream (20 mL/minutes) in a flow reactor
system with feed/catalyst ratio of 5. The activity tests were operated at the various temperatures, i.e.: 350 °C, 400 °C,
and 450 °C and the various catalysts, i.e.: Ni/Z, Ni-Mo/Z, Co/Z, and Co-Mo/Z to determine the optimum conditions
for hydrocracking of LDPE plastics. The conversion yield could be calculated by using the following equations:
218 Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224

ௐ೗
™–Ψ‘ˆŽ‹“—‹†›‹‡Ž† ൌ šͳͲͲΨ
ௐ೑
(1)
ܹ௦ଶ െ ܹ௦ଵ
™–Ψ‘ˆ•‘Ž‹†›‹‡Ž†ൌ šͳͲͲΨ

(2)
ܹ௥
™–Ψ‘ˆ”‡•‹†—‡ൌ šͳͲͲΨ
ˆ (3)
™–Ψ‘ˆ‰ƒ•›‹‡Ž†ൌͳͲͲΨ െ ™–ΨሾŽ‹“—‹†›‹‡Ž† ൅ •‘Ž‹†›‹‡Ž† ൅ ”‡•‹†—‡ሿ
(4)
x Wl = the weight of liquid yield
x Ws1 = the weight of catalyst sample before hydrocracking process
x Ws2 = the weight of catalyst sample after hydrocracking process
x Wr = the weight of residue of feed
x Wf = the weight of feed

The liquid yields were analyzed by Gas Chromatography (GC) using HP-5890 GC Packard Series II, CBP-1
column (cross-linked 95% dimethyl 5% diphenylpolysiloxane), length 30 m, diameter 0.25 mm, thickness 0.50 μm,
FID detector, temperature max.325 °C;Gas Chromatography–Mass Spectrometry (GC-MS QP2010S Shimadzu) and
using AGILENTJ&W DB-1 column (crossbond 100% dimethylpolysiloxane), length 30 m, diameter 0.25 mm,
thickness 0.25 μm, and temperature 60 – 325 °C.

3. Results and discussion

3.1. Catalysts characterization

Zeolites are crystalline aluminosilicates whose main feature is the occurrence of a microporous structure. They
are formed by the linkage of SiO44- and AlO45- [7,8]. Zeolites show a monodimensional, bidimensional, or
tridimensional pore channel system with pore sizes in the range of 0.4 – 1.0 nm, which turns them into molecular
sieves (shape selectivity) [11].

Wave number (cm-1)

Fig.1.Infrared spectrum of natural zeolite: (a) before acid treatment(Zo); (b) after acid treatment (Z).
Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224 219

M
M = Mordenite
M M M M

E
M
M M M M
D
Intensity (a.u)

M
M M M
M

C
M
M M
M M

B
M
M
M
M M

0 10 20 30 40 50 60 70
2θ (deg)

Fig. 2. XRD patterns of (a) activated natural zeolite (Z); and loaded metal (b) Ni/Z; (c) Ni-Mo/Z; (d) Co/Z; (e) Co-Mo/Z.

The treatment of natural zeolite was capable to eliminate the mineral impurities and reduce the amount of
aluminum (Al) in the framework (dealumination) so it could produce the sites that have high acid strength. Fig. 1
shows that the acid treatment caused the dealumination of natural zeolite sample, indicated by the shifted
asymmetric stretching vibration of Si-O-Al of wave number from 1041.56 cm-1 to 1080.14 cm-1. The shift towards
larger wave number was due to the fact that the decreasing of Al framework becomes Al non framework
(dealumination). Consequently, the ratio of Si/Al increased. It led to be more hydrophobic and had greater thermal
stability ascatalyst in the hydrocracking reaction. The activity of zeolite was measured by the amount of water
adsorbed in the zeolite framework as well. Water could poison the active sites on zeolite. After the acid treatment,
water intensity at 3464.15 cm-1 (stretching vibration of O-H) and 1635.64 cm-1 (bending vibration of O-H) decreased.
It explains that zeolite pores were more opened and clean from impurities.
X-ray diffraction (XRD) was applied to identify the type of zeolite mineral contained in the sample and the
crystallinity of zeolite sample. The type of mineral was shown by the position of diffraction angle (2θ (degree)),
while the crystallinity of a sample was shown on its intensity. Fig. 2 shows the XRD pattern of zeolite samples. It
was identified that the type of natural zeolite mineral obtained from Sukabumi, West Java Province of Indonesia
consisted of mordenite type crystalline. Fig. 2 shows five main peaks in the 2θ region, i.e.: 9.828°, 22.392°, 26.440°,
26.640°, and 28.016°; they were the characteristic of mordenite according to JCPDS (Joint Committee on Powder
Diffraction Standards) data. Loading metals onto natural zeolite led to the decrease crystallinity as indicated by a
decrease in the intensity of some characteristic peaks. Loading of bimetal did not affect the crystallinity when
compared with the crystallinity of monometal catalyst. Overall, the treatment of zeolite and metals impregnation
onto zeolite did not ruin the crystallinity significantly.
The catalysts are usually employed for hydrocracking of plastic containing Lewis or Brönsted acid sites or a
combination of both of them. In this study, the determination of acid sites using ammonia base vapour as adsorbate
gave the total amount of acid sites with the assumption that the small size of ammonia molecule made it possible to
enter the pores of the catalyst sample. The amount of acid sites using pyridine base vapour as adsorbate was the
amount of acid sites on its surface with the assumption that the relatively large size of pyridine molecule could only
interact with the acid sites on the outer surface. The amount of acid sites of each catalyst samples is presented in
Table 1.
220 Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224

Table 1.The amount of acidity on catalyst samples before and after acid treatment and metals impregnation.
Catalyst samples Ni/Co/Mo ratio Acid amount (mmol/g)
Total Surface Pores
Zo 0:0:0 1.003 0.187 0.816
Z 0:0:0 1.751 0.228 1.523
Ni/Z 1:0:0 3.420 0.394 3.026
Ni-Mo/Z 1:0:1 2.950 0.215 2.735
Co/Z 0:1:0 4.129 0.583 3.546
Co-Mo/Z 0:1:1 3.241 0.439 2.802

Table 1 shows that metal loaded on zeolite increased the acid amount significantly either using ammonia or
pyridine base vapour as adsorbate. It presented that the acid amount was affected by the size of transition metal. The
highest acid amount for monometallic catalyst was reached by Co/Z,3.546 mmol/g in the poresand 0.583 mmol/gon
the surface. The highest acid amount for bimetallic catalyst was reached by Co-Mo/Z,2.802 mmol/g in the poresand
0.439 mmol/g on the surface. Based on a journal by Datka et al. [9], the acid sites on zeolite consist of Brönsted and
Lewis acid. The acid treatment caused the zeolite pores more opened and zeolite surface wider so that many
Brönsted acid sites formed were consequently more effective in ammonia base vapour adsorption.
The determination of acidity for each catalyst had the same pattern, either using ammonia or pyridine base
vapour as its adsorbate. Generally, the catalyst acidity of natural zeolite was increased significantly by loaded
monometal or bimetal. This could be understood from metal characteristic of Ni, Ni-Mo, Co and Co-Mo which was
dispersed on the surface and in the pores of zeolite. The empty d-orbital or half-filled of metal could accept the
electron pair from adsorbate effectively. Contribution of acid sites on the Ni, Ni-Mo, Co, and Co-Mo metals was
Lewis acid sites. Hegedus et al. [10] suggested that the surface of metal oxide catalyst such as zeolite consists of
oxygen atom, hydroxil group, and some unknown metal atom. Chemical characteristic of this species interacted with
loaded metal which is the strength affected by localized electron. Anion of oxygen acts as Lewis base, metal acts as
Lewis acid, and hydroxil group acts as acid or base. The strength and concentration of acid and base sites depend on
metal-oxygen bond (M-O). Acid oxide has covalent bond, while base oxide has ionic bond.

3.2. Hydrocracking experiments

This experiment was performed initially a pyrolysis step of the solid plastic wastes feed, followed by a catalytic
hydrocracking of the heavy oil obtained previously with a view to further optimize the properties of the obtained
final hydrocarbon mixture. Thus, the viscosity of the feed decreased and the impurities might be removed by
different treatments [11]. Fig.3 shows the activity of Ni/Z catalyst on the hydrocracking of LDPE plastics at the
various temperatures, i.e.: 350 °C, 400 °C, and 450 °C in a flow reactor system with feed/catalyst ratio of 5.

90
Gas
80 75.18 74.21 Liquid
70 68.27
Solid
60
Yield (wt %)

50
40
30.64
30 23.89 24.69
20
10 0.94 1.10 1.10
0
350 400 450
Temperature (°C)

Fig. 3.The conversion yield from LDPE plastic waste at various temperatures over Ni/Z catalyst
Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224 221

90 86.30 Gas
83.71
80 75.18 76.00 Liquid
70 Solid
Yield (wt %) 60
50
40
30 23.88 23.92
20 12.20 14.91
10 0.94 0.92 1.39
1.51
0
Ni/Z NiMo/Z Co/Z CoMo/Z
Catalyst
Fig. 4. The conversion yield from LDPE plastic waste over all various catalyst samples at 350 °C.

According to the figure above, it was observed that the conversion produced the gas yield, liquid yield, and solid
yield with the total conversion of 100wt%. There was no remaining residue. The liquid yield was clear yellowish
like-gasoline. The solid yield was black coke formed on catalyst surface. In Fig.3, it can be clearly observed that the
hydrocracking process over Ni/Z catalyst produced the highest liquid yield at 450 °C, while the least coke was
obtained at 350 °C. Increase in LDPE plastic conversion could be observed when temperature was increased from
350 to 450 °C, possibly due to the catalyst activity enhanced with temperature above 350 °C. According to the
Arrhenius equation, reaction rate constant increases with temperature. Hence, the molecules are able to interact more
with each other to produce the liquid yield [14]. However, the conversion at the various temperatures showed
relatively similar percentage of liquid yield as the desired product, so it was concluded that the most effective
temperature for hydrocracking of LDPE plastics over Ni/Z catalyst was at 350 °C, with the consideration of the
lower temperature so the operating cost would be smaller.
The hydrocracking of LDPE plastics over the various catalysts at 350 °C is described at Fig.4. All catalysts were
able to convert completely the plastic feed into gaseous and liquid hydrocarbons, showing low coking levels. The
highest liquid yield was produced over Ni/Z catalyst, while the lowest coke was produced over Co/Z catalyst.
Zeolite based catalyst has a very narrow pore size distribution in the micropore range and possess strong acidity. It is
reasonable to assume that the polymer macromolecules of plastics break down first on the external catalytic surface,
and only small fragments can enter the catalyst pore structure to undergo further reactions. With zeolites, this
fragmentation progresses to quite small molecules and increases the yield to gaseous hydrocarbons, which are
considered inferior to liquid fuel [12].

80
71.49
70 68.40 67.60 Gasoline
64.90
Diesel
60
Selectivity (wt %)

50
40 35.10 32.40
31.60
28.52
30
20
10
0
Ni/Z NiMo/Z Co/Z CoMo/Z
Catalyst

Fig. 5. The selectivity of liquid yield from hydrocracking process over allvarious catalyst samples
222 Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224

Fig.5 shows the product distribution of liquid fraction that consisted of the gasoline (C 5 – C12) in retention time
range of 0 – 10 minutes, as well as the diesel (C13 – C20) in retention time range of > 10 minutes based on
chromatogram of GC[6]. The overall hydrocracking was performed the higher gasoline product than diesel product.
The highest conversion was shown by Co-Mo/Z catalyst. It has been mentioned previously that the acid amount of
catalyst is closely related to its activity for the hydrocracking process [13]. This tendency almost appeared on this
study. According to the previous acidity test in Table 1, Co/Z and Co-Mo/Z catalysts had the higher acid amount
than Ni/Z and Ni-Mo/Z catalysts. In other words, when the acidity of catalyst increased, the desirable liquid product
of conversion like gasoline increased as well.
The catalytic effects of these catalysts in hydrocracking were due to their acidic properties which have been
explained by a carbonium ion theory. The zeolite based catalyst consisted of acid sites that was able to hydrocrack
the LDPE fed via the carbonium ion intermediate. Initiation might occur on some defect sites of the LDPE chains.
For instance, an olefinic linkage could be converted into an on-chain carbonium ion by proton addition. Then the
LDPE chain might be broken up through β-scission. Initiation might also take place through random hydride-ion
abstraction by low-molecular-weight carbonium ions. The molecular weight of the main polymer chains could be
reduced through successive attacks by acidic sites or other carbonium ions and chain cleavage, and yielding an
oligomer fraction (approximately C30 – C80). Further cleavage of the oligomer fraction which is probably by direct β-
scission of chain-end carbonium ions leads to gas fraction(approximately C1 – C4) and a liquid fraction
(approximately C5 – C20). The carbonium ion intermediates could undergo rearrangement by hydrogen or carbon
atom shifts leading to, e.g. a double-bond isomerization of an olefin. Other important isomerization reactions are
methyl-group shift and isomerization of saturated hydrocarbons. Some carbonium ion intermediates could undergo
cyclization reactions. An example is when hydride-ion abstraction first takes place on an olefin at a position several
carbons removed from the double bond, the result is the formation of an olefiniccarbonium ion. This carbonium ion
could undergo intramolecular attack on the double bond [15]. Through this mechanism, it was produced branced
alkane (isoparaffin), n-alkene (olefin), and cycloalkane (naphthene) in the range of liquid hydrocarbons with the
number of carbon atom between C6until C19. The distribution of carbon atom is shown at Fig. 6 below.
Based on GCMS data, conversion of LDPE plastics into hydrocarbon compounds with atom chain C 6 until C19
indicated the fuel chemical structure that could be observed at Fig.7. The presence of these sorts of hydrocarbons
was positive for the formulation of gasoline to increase in the research octane number (RON) values. It was
favorable for diesel as well, to improve the pour and cloud point. On Fig.7, there were the structures of isoparaffins,
naphthenes, and olefins. They were the characteristic chemical compounds of fuel. Olefins were the dominant
products in this hydrocracking experiment, and apparently in this system, they were not suffering important
hydrogen transfer that saturated them into paraffins. For further studies in this field using these catalysts, it was
supposed to improve the techniques like secondary reactions that took place extensively, such as isomerization and
aromatization.

40

35

30
Selectivity (wt%)

25

20

15

10

0
C6 C7 C8 C9 C10 C11 C12 C13 C16 C17 C19
The number of carbon atom

Fig. 6. The distribution of carbon atom number of liquid yield from hydrocracking process over Ni-Mo/Z
Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224 223

Olefins 58.07

Naphthenes 14.12

Isoparaffins 4.97

0 10 20 30 40 50 60 70
Selectivity (wt%)

Fig. 7. The GCMS characterization data of liquid yieldfrom hydrocracking process over Ni-Mo/Z.

The presence of catalyst on the hydrocracking process lowered both the apparent activation energy of the plastic
degradation and the frequency factor with regard to pure thermal cracking. However, the effect of the decrease in the
activation energy was higher than that of the frequency factor, so the final result was an enhancement of the reaction
rate. Thus, the output of gases and liquid distillates collected after reaction was increased. This final result
established the possibility of LDPE plastic waste as fuel through hydrocracking process over metals-supported on
natural zeolite.

4. Conclusions

From the result above, it has been found the type of natural zeolite mineral obtained from Sukabumi, West Java
Province of Indonesia consisted of the mordenite type crystalline that have a high thermal stability. Acid treatment
and metals impregnation onto natural zeolite as support material of monometal and bimetal catalysts could increase
the activity and selectivityfor hydrocracking processes, due to the increased acid amount. The result of
hydrocracking experiment clearly shows that fuel can be produced from LDPE plastic waste over all various bi-
functional catalyst, either monometal or bimetal. The presence of metal loaded on natural zeolite was proved for
conversion of LDPE plastic waste into gasoline and diesel fraction with increasing of acid amount, indicated by Co-
Mo/Z catalyst that has the highest gasoline conversion. Hydrocracking of LDPE plastic waste over Co-Mo/Z at
350 °Ceffectively produced 14.91% of liquid yield, 1.39% of solid yield and 83.71% of gas yield. From the liquid
yield, it was estimated as much as 71.49% of gasoline range (C 5 – C12) and 28.52% of diesel range (C13 – C20). The
liquid yield consisted of hydrocarbon compounds such as isoparaffins, olefins, and naphthenes with atom chain
between C6 until C19. The gas yield could be short-chain hydrocarbons (C1 – C4) that can be used as gas fuel.

Acknowledgements

This work was financially supported by Department of Chemistry, Faculty of Mathematics and Natural
Sciences,Universitas Gadjah Mada, Yogyakarta, Indonesia (Project No. 0031/J01.1.28/PL.06.02/2013).The authors
also gratefully acknowledge the funding support from Karya Salemba Empat Foundation, Indonesia.
224 Wiwin Sriningsih et al. / Procedia Environmental Sciences 20 (2014) 215 – 224

References

1. Aguado J, Serrano DP, Escola JM.Fuels from waste plastics by thermal and catalytic processes: a review.Ind. Eng. Chem. Res.2008;47: 7982-
7992.
2. Serrano DP.The current status of plastics recycling in Europe.Proceedings of ISFR (2007) September 16–20, Jeju, Korea.
3. Ding W, Liang J, Anderson LL.Hydrocracking and hydroisomerization of high-density polyethylene and waste plastic over zeolite and silica-
alumina-supported Ni and Ni-Mo sulfides.Energy & Fuels 1997;11:1219-1224.
4. Uemichi Y, Nakamura J, Itoh T, Sugioka M, Garforth AA, Dwyer J. Conversion of polyethylene into gasoline-range fuels by two-stage
catalytic degradation using silica-alumina and HZSM-5 zeolite. Ind. Eng. Chem. Res.1999;38:385-390.
5. Escola JM, Aguado J, Serrano DP, Briones L, Díaz de Tuesta JL, Calvo R, Fernandez E. Conversion of polyethylene into transportation fuels
by the combination of thermal cracking and catalytic hydroreforming over Ni-supported hierarchical beta zeolite.Energy Fuels
2012;26:3187-3195.
6. Trisunaryanti W, Triyono, Rizki CN, Saptoadi H, Alimuddin Z, Syamsiro M, Yoshikawa K. Characteristics of metal supported-zeolite
catalysts for hydrocracking of polyethylene terephthalate.IORS–JAC2013;3:29-34.
7. Bhatia S.Zeolite catalytisis: principles and applications. FL: CRC Press Inc.; 1990.
8. Oudejans JS.Zeolites catalyst in some organic reactions.1st ed. Holland: Chemical Research; 1984.
9. Datka J, TurekAM, Jehng JM, Wachs IE. Acidic properties of supported niobium oxide catalyts: an infrared spectroscopy investigation.J.
Catal. 1992;135:186-199.
10. Hegedus LL, Aris R, Bell AT, Boudart M, Chen NY, Gates BC, Haag WO, Somorjai GA, Wei J.Catalyst design progress and
prospective.New York: John Wiley & Son; 1987.
11. Serrano DP, Aguado J,Escola JM. Developing advanced catalysts for the conversion of polyolefinic waste plastics into fuels and
chemicals.ACS Catal. 2012;2:1924-1941.
12. Manos G, Yusof IY, Gangas NH,Papayannakos N. Tertiary recycling of polyethylene to hydrocarbon fuel by catalytic cracking over
aluminum pillared clays.Energy & Fuels 2002;16:485-489.
13. Chang N,Gu Z, Wang Z, Liu Z,Hou X, Wang J.Study of Y zeolite catalysts for coal tar hydro-crackingin supercritical gasoline. J Porous
Mater 2011;18:589-596.
14. Yu Z, Chen L, Zong Z, Zhu Z, Wu Q.The effects of temperature and hydrogen partial pressure onhydrocracking of
phenanthrene.International Journal of Chemistry2011; 3:67-73.
15. BuekensAG, Huang H. Catalytic plastics cracking for recovery ofgasoline-range hydrocarbons from municipalplastic wastes. Resources,
Conservation and Recycling 1998;23:163-181.
16. de la Puente G, Klocker C, Sedran UǤ Conversion of waste plastics into fuels:Recycling polyethylene in FCC. Applied Catalysis B:
Environmental2002;36:279-285.
17. Kaneda K, Wada T, Murata S, Nomura M. Hydrocracking of dibenzothiophenes catalyzed bypalladium- and nickel-coloaded Y-type zeolite.
Energy & Fuels 1998;12:298-303.
18. DastanianM, Azad FS. Desulfurization of gasoline over nanoporous nickel-loaded Y-type zeolite atambient conditions. Ind. Eng. Chem. Res.
2010;49:11254-11259.
19. de Haan R, Joorst G, Mokoena E, Nicolaides CP. Non-sulfided nickel supported on silicated alumina as catalyst for the hydrocracking of n-
hexadecane and of iron-based Fischer-Tropsch wax. Applied Catalysis A: General 2007;327: 247-254.
20. Hwang SH, Lee JH, Park SY, Park DR, Jung JC, Lee SB, Song IK. Production of middle distillate through hydrocracking of paraffin wax
over NiMo/SiO2–Al2O3 catalysts: effect of SiO2–Al2O3 composition on acid property and catalytic performance of NiMo/ SiO2–Al2O3
catalysts. Catal Lett 2009;129:163-169.

Вам также может понравиться