You are on page 1of 15

JOEI479_proof ■ 13 August 2018 ■ 1/15

Journal of the Energy Institute xxx (2018) 1e15

Contents lists available at ScienceDirect

Journal of the Energy Institute


journal homepage: http://www.journals.elsevier.com/journal-of-the-energy-
institute

1 Co-pyrolysis of waste polypropylene and rice bran wax‒ production of


2
3 biofuel and its characterization
4
5 Q4 Akancha, Namrata Kumari*, R.K. Singh
6
Department of Chemical Engineering, National Institute of Technology, Rourkela 769008, India
7
8
9
10 a r t i c l e i n f o a b s t r a c t
11
Article history: The present study aims to investigate the interaction during co-pyrolysis of Polypropylene (PP) and Rice
12 Received 20 March 2018 bran wax (RBW). Initial characterization of feedstock was found to be suitable to carry out further
13 Received in revised form experimental sets. Further, the pyrolysis experiments of PP, RBW and different blends (1:1, 1:2, 1:3, 2:1
14 27 July 2018
and 3:1) were carried out in a semi-batch reactor. As per TGA analysis the temperature range between
15 Accepted 31 July 2018
400  C and 650  C at a constant heating rate of 25  C/min was determined. The maximum liquid yield of
Available online xxx
16 PP, RBW was approximately 76%, 86% at the temperature of 500  C and 600  C respectively. Whereas
17 maximum liquid yield from co‒pyrolysis was obtained at 1:3 blend i.e. 81% at 550  C.
Keywords:
18 Co-pyrolysis GC‒MS results inferred the highest percentage of hydrocarbon whereas 1:3 blends has lower oxygen
19 Polypropylene containing groups than RBW in liquid products. FTIR data of all blends indicates higher range of alkyl and
20 Rice bran wax aromatic compounds. H1 NMR results also confirmed the higher compounds into aliphatic region than
21 Pyrolytic liquid aromatics or heteroaromatics groups. Further, most of the fuel properties of 1:3 blend falls within the
1:3 blend range of gasoline and diesel properties. Study was extended to know crystallization behaviour of fuel by
22
DSC analysis from two consecutive heating and cooling cycles of 50 to 60  C and reversed till 50  C at
23
10  C min 1. Two peaks at 24  C and 26  C were observed during heating cycle whereas single peak at
24
23  C during cooling cycle. 1:3 blend residual char characterization was also included in the work. Un-
25 fortunately, the SEM and BET results inferred that the char was not highly porous.
26 © 2018 Energy Institute. Published by Elsevier Ltd. All rights reserved.
27
28
29
30
31 1. Introduction
32
33 Fossil fuel contributed significantly for growth of industrialization and economy worldwide. Although precise data available in public
34 domain is rather contradictory but the era of cheap petroleum fuel is certainly over. According to IEA, 2008 report, it was also expected that
35 almost all additional demand will arise maximum from China and India [1,2]. However, supply and demand equilibrium could only be
36 balanced by developing unconventional resources. Pursuing energy security policies could not just be economic but also have to secure
37 environmental pollution barriers.
38 In account of exploring new alternative energy resources, biofuel is receiving importance in recent research. But suitable source of biofuel
39 is still topic of consideration [2]. However, RBW is a natural wax and residual derivative from RBO refining industry. It has been used in the
40 cosmetic, pharmaceutical, food, polymer, leather industries and so on. Rice is most overwhelming cereal, grown in more than 100 countries
41 and about 25% of the world's food grain production alone [3]. Recent studies declaims that the RBO is unfit for human consumption due to
42 several factors. Therefore, RBW and RBO has been utilized for biofuel production in studies [4,5]. Pyrolysis experiment results of RBW
43 confirms presence of alkane, alkene; alkane, olefins as main component in liquid and gas products respectively [6].
44 Natural waxes represents group of several complex structured lipids which could be grouped in fatty acids, fatty alcohols and hydro-
45 carbons. Potential use of wax depends upon identification of wax profiles therefore, existing literature mainly focus on analytical identi-
46 fication of natural wax materials. The PyeGCeMS was used to analyse bleached beeswax, lanolin and yellow carnauba wax [7]. Similarly,
47 bulk chemical profile generation of RBW, sunflower wax, bees wax, candelilla wax, carnauba Brazilian wax, carnauba wild wax, berry wax
48
49
50
51 Q1 * Corresponding author. Q2
E-mail address: namratajaiswal@gmail.com (N. Kumari).
52
53
https://doi.org/10.1016/j.joei.2018.07.011
54 1743-9671/© 2018 Energy Institute. Published by Elsevier Ltd. All rights reserved.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 2/15

2 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1 were determined by HPLCeELSD and GCeMS [8]. However, less attempts have been made on using natural wax as fuel by pyrolysis method.
2 Even though, properties of RBW was studied but data on using it as pyrolytic oil is rare [6].
3 Co-pyrolysis of biomass with waste plastics greatly influence the yield, the chemical structure and the physical properties of final
4 products. These polymers donates hydrogen during thermocatalytic conversion and therefore increase the liquid production [9]. Therefore,
5 several studies have been evaluated on co‒pyrolysis of polyolefins with biomass. The synergistic effect and rise in liquid product by addition
6 of plastics (PE, PP, PS) with pine sawdust and waste paper has been claimed [1,10]. Further, co-pyrolysis of cellulose, lignin, saw dust, rice
7 husk and corn cob with PP results in hydroxyl radical shifts from the cellulose chain to polypropylene radicals to produce long chain alcohols
8 with high carbon numbers [11e13]. PP blend with LDPE shows high volatile hydrocarbon conversion with high PP in blend [14]. Even, co‒
9 pyrolysis of PP with Soma-lignites form Turkey resulted in lower amount of coke deposit in the presence of PP. Studies also suggests that
10 bond dissociation energy of PP than PE results in higher co‒liquefaction yield at different temperatures with or without a catalyst [15]. PP is a
11 thermoplastic polymer and widely utilized in packaging, textiles, automotive parts, equipment and reusable containers. Therefore, it
12 contributes a key role of non‒biodegradable plastic waste that needs to be recycled and recovered [16].
13 Previously, numerous research work has been carried out by various researches to characterize wax and pyrolytic liquid from blends of
14 biomass with PP (Table 2). Therefore, for the first time our study reports the physical and chemical characterization of liquid product
15 obtained along with bio-char from co-pyrolysis of RBW and PP. Experiments were carried out in a semi-batch reactor in the temperature
16 range of 300 Ce650  C with varying RBW and PP ratios of 1:1, 1:2, 1:3, 2:1 and 3:1 at a heating rate of 20 ± 1  C/min. Co‒pyrolysis of RBW/PP
17 was investigated with the aim of upgrading the oil yield and quality improvement. Also, all the abbreviated words used throughout the
18 manuscript is lists in Table 1.
19
20
2. Experimental
21
22
2.1. Materials
23
24
Waste PP (used plastic disposable glasses) were collected from the Lecture Annexure, National Institute of Technology Rourkela campus
25
waste disposal courtyard and were cut into flakes by a shredder. RBW was collected from a rice mill in Raipur, Chhattisgarh. The waste PP
26
and RBX samples were characterized for melting point by the DSC (NETZSCH DSC 200F3); calorific value by the bomb calorimeter (Parr 6100
27
Digital Bomb Calorimeter); moisture, volatile matter, ash, fixed carbon by ASTM D3173-75 and Ultimate analysis by elemental analyser
28
(VARIOEL CHNS).
29
30
31 2.2. Pyrolysis experimental setup
32
33 TGA of PP, RBW and 1:1 (PP: RBW) raw material was done using the TOSHVIN model DTG-60 Series. A known weight of the sample was
34 taken in a platinum crucible and at 10  C/min heating rate. The inert environment was created by stream of pure nitrogen gas with a flow
35 rate of 50 ml/min from 25  C to 800  C.
36 As per TGA analysis, further, pyrolysis was done to find the optimum temperature of the maximum liquid yield. A total of 20 g of the
37 feedstock was taken for thermal pyrolysis at temperature range of 300 Ce650  C in steps of 25  C and at a constant heating rate of
38 20  C/min. Individual feedstock and mixtures of PP and RBW with weight ratios of 1:1, 2:1, 3:1, 1:2 and 1:3 were adopted for pyrolysis.
39
40
Table 1
41 Common abbreviation and nomenclature used in this manuscript.
42
Abbreviation/Nomenclature Description
43
44 ABS Acrylonitrile Butadiene Styrene
45 BET Brunauer, Emmett and Teller
DSC differential scanning calorimetry
46
FTIR Fourier transformation infrared spectroscopy
47 GC‒MS Gas chromatography‒mass spectroscopy
48 GCV Gross Calorific Value
49 H1 NMR Proton nuclear magnetic resonance spectroscopy
HDPE High density polyethylene
50
HPLCeELSD High performance liquid chromatography‒evaporative light scattering detector
51 LDPE low density polyethylene
52 PA Polyamide
53 PBT Polybutylene Terephthalate
54 PC Polycarbonate
PE polyethylene
55
PET polyethylene terephthalate
56 POM Polyacetal
57 PP Polypropylene
58 PS polystyrene
PVC Polyvinly Chlorine
59
Py‒GC Pyrolysis gas chromatography
60 RBO rice bran oil
61 RBW Rice bran wax
62 RH rice husk
63 SEM Scanning Electron Microscopy
TGA Thermogravimetric analysis
64
TMS tandem mass spectrometry
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 3/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 3

1 Table 2
2 Comparison studies of product yield from different feedstocks.

3 Feedstock Conversion technique Pyrolysis Temperature Reference


4 Rice bran Soxhlet extraction (hexane) followed by Acid-catalyzed [4]
5 methanolysis
6 RBW, sunflower wax, bees wax, candelilla wax, carnauba solvent extraction by chloroform/hexane [8]
7 wax, berry wax

beeswax, lanolin, carnauba wax PyeGCeMS 800 C [7]
8
RBW Conventional Pyrolysis 600  C [6]
9 PyeGCeMS 600  C, 800  C
10 RH fast pyrolysis in electrothermal fluidized-bed 500  C [17]
11 RH Catalytic pyrolysis in fixed-bed reactor 550  C [18]
12 RH, sawdust fluidized-bed reactor 420  C, 540  C [19]
Rice bran, HDPE Co-pyrolysis by thermo-gravimetric analyses (TGA) ambient temperature to 900  C [9]
13 HDPE, PP, PET, Paper Catalytic co-pyrolysis in fixed bed reactor 800  C [10]
14 PE, PP, PS, pine sawdust catalytic co-pyrolysis in continuous feeding fluidized-bed 450 C‒650  C [1]
15 reactor
16 cellulose, xylan, lignin, PP co-pyrolysis in fixed bed 450  C [11]
reactor
17
RH, corn cob, sawdust, PP co-pyrolysis in fixed bed 450 C‒600  C [13]
18 reactor
19 Cellulose, PP TGA 
300e375 C, 400e500 C 
[12]
20 LDPE, PP co-pyrolysis in fixed bed 300e600  C [14]
21 reactor
Turkish lignite, PP co-pyrolysis in fixed bed 300e600  C [15]
22 reactor
23
24
25
26 The parameters such as reaction time, liquid yield, solid residual yield and vapour was calculated by difference during experiment. The
27 variation of these parameters was plotted against time.
28 The pyrolysis setup used in the work was semi‒batch reactor of volume 300 mL (Fig. 1). Briefly, the stainless steel reactor sealed at one
29 end and an outlet tube at another end for obtaining the various products of the reaction. The cylindrical reactor was having dimension of
30 about 5150 mm of height, 540 mm of internal diameter and 544 mm of external diameter. It was heated externally by an electric furnace. The
31 temperature sensitivity was measured by a CreAl: K type thermocouple which was controlled by external PID controller. Continuous ni-
32 trogen flow at the rate of 5 mL/min was maintained during all the experiments. Reaction vapors were condensed in the glass condenser by
33 circulating continuous water flow and the produced liquid was collected in a flask.
34
35 2.3. Characterization of pyrolytic liquid product
36
37 The pyrolytic oil has been characterized for physical and chemical properties. Physical properties such as specific gravity, density, ki-
38 nematic viscosity, conradson carbon residue, flash point, fire point, pour point, gross calorific value of the pyrolytic oil was determined using
39 the Indian Standard methods i.e. I.S.1448 P.16, I.S.1448 P.16, I.S.1448 P.25, I.S:1448 P:122, I.S.1448: P:20, I.S:1448 P:20, I.S.1448 P:10 and I.S.
40 1448: P:6 respectively. The density and specific gravity measurement were done with an accuracy of ±0.0005 gm/cc and the other
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Fig. 1. Schematic representation of experimental setup. A. Nitrogen source, B. PID controller, C. Electrically heated furnace, D. stainless steel reactor, E. feed, F. thermocouple,
65 G. condenser, H. water circulation, I. pyrolytic liquid collection, J. clamp.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 4/15

4 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1 parameters such as pour point, flash point and fire point was measured with ±1  C accuracy. The mean value was calculated by repeating
2 every experiment thrice and results were considered by one-way analysis of variance (ANOVA). The cold flow properties of the pyrolytic
3 liquid phases were determined by DSC (NETZSCH DSC 200F3). Five milligrams of the pyrolytic liquid was added to a 40 ml sealed pan under a
4 nitrogen flow of 50 mL min1. A temperature program with two consecutive dynamic steps, in which the sample was first heated from 50
5 to 60  C followed by a second cooling step at 10  C min 1 to 50  C at 10  C min 1 was employed [20].
6 Chemical properties of pyrolytic oil was determined by FTIR, GC‒MS and NMR. The FTIR spectra were collected in the range of
7 400e4000 cm1 with 8 cm1 resolution by using Perkin Elmer RX. The GC‒MS analysis was carried out to know the composition of liquid
8 derived from co‒pyrolysis of RBW and PP which was analyzed using 7890B GC System/5977A MSD Agilent. The programme followed was at
9 70  C held for 3 min and ramped to 200  C (at 10  C/min) for the total run time of 32 min. The DB-5 ms column of diameter 0.25 mm and
10 30 m length was used. The volume of 1 mL liquid product was injected into the column with the carrier gas (helium) at 1.5 mL/min flow rate.
11 The compounds of liquid product were ionized at 70 eV ionization energy. The ion sources temperature was 200  C and were analyzed over a
12 mass electron (m/z) range of 40e700. The chromatogram obtained at different retention time, and respective mass spectra were plotted and
13 compared with the spectral data of the NIST library.
14 The 1H NMR spectra were recorded by using a 400 MHz, BRUKER DPX-400, High-performance digital FT‒NMR spectrometer where
15 chloroform-d containing tetramethylsilane used as the internal standard.
16
17 2.4. Characterization of pyrolytic char
18
19 The physical, morphologic, crystalline or powdered material characteristics of the bio‒char was studied by proximate analysis, elemental
20 analyser, GCV, SEM and BET analysis. The char product obtained from the PP: RBW ratio pyrolysis was characterized by SEM (Model: NOVA
21 NANO SEM450_CR_NITRKL) at different magnification values to observe a clear view on porosity. The surface area of char was estimated by
22 nitrogen absorption at 77.35 K using an automatic adsorption instrument (Quantachrome® ASiQwin™-/Instrument name-: Autosorb iQ
23 Station 1). Before finding the gas adsorption measurement, the char was outgassed in the vacuum chamber at 250  C for a time of 5 h to
24 remove the moisture and other impurities that were sticked to the surface of the char sample.
25
26
3. Results and discussion
27
28
3.1. Characterization of raw materials
29
30
The important characterization parameters were included in this study to know the suitability of raw material to be used as fuel. It
31 includes the detail analysis of melting point by DSC, GCV, ultimate analysis, proximate analysis and has been presented in Table 3. The
32
melting point specification of PP and RBW were found to be similar from other studied literature [8,12]. The obtained moisture content,
33
volatile matter, ash content and fixed carbon of PP were 0.67%, 98.64%, 0.24% and 0.45% respectively. Whereas the value obtained for similar
34
parameters of RBW were 5.45%, 90.41%, 1.54% and 2.6%. Obtained results for PP were comparable with available proximate study but
35
literature for RBW is rare [6]. High volatile matter, low ash and moisture content of PP and RBW favoured the high bio‒oil yield during
36
pyrolysis reaction. The elemental analysis shows the higher amount of carbon, hydrogen as compared to oxygen, nitrogen and sulfur in both
37
raw materials. Similarly, high calorific value of both the selected feedstock suggest satisfactory result for bio‒oil production.
38
Feedstock characterization is basic estimation for understanding the impact on biodiesel production. All the biomass could not be
39
suitable, qualitative, economic and high yield bio‒oil producer. Usually, feedstock characterization has been underestimated in studies but
40
could be significant during business planning. Volatile matter % of PP was reported to be higher than other electronic plastic waste i.e. PS,
41
PVC, PBT, PC and ABS. Further, carbon content of PP was found to be higher than ABS, PA, PC, LDPE, ABS, PBT, PVC and POM [21]. Similarly,
42
Zhou et al., 2013 reported the GCV of rice husk as 17.21 MJ/kg; volatiles as 71.47 wt % and fixed carbon as 18.39 wt % which are very low in
43
comparison to RBW content [18].
44
45
3.2. Decomposition and product distribution
46
47
TGA is widely applicable for feedstock analysis to estimate thermal stability, oxidation or combustion reaction, reaction kinetics, acti-
48
vation energies and could be coupled with other techniques such as FTIR or GC‒MS to know the detailed changes with respect to
49
50
51 Table 3
52 Characterization of Polypropylene, Rice Bran Wax and obtained residual char of blend (1:3).
53 Properties Polypropylene Rice Bran Wax Char (1:3)
54  
Melting point 166 C 79 C e
55
Proximate Analysis (wt. %)
56 Moisture content 0.67 1.45 1.57
57 Volatile matter 98.64 91.49 27.8
58 Ash content 0.24 1.74 19.3
Fixed carbon 0.45 5.32 51.33
59
Ultimate Analysis (wt. %)
60 Carbon (C) 83.74 76.76 92.6
61 Hydrogen (H) 14.27 15.13 1.36
62 Nitrogen (N) 0.01 0.02 0.04
63 Sulphur (S) 0.84 0.68 0.83
Oxygen (O) (By difference) 1.14 7.41 5.17
64
GCV (MJ/Kg) 46.44 40.65 27.26
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 5/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 5

1 temperature/time [22]. Liu et al., 2011 reported that crystalline nanotubes formation were accelerated at high decomposition temperature
2 which was identified by TGA and XRD analysis [16]. In present work, the thermal degradation of samples was carried out in order to know
3 the thermal properties and the weight loss with respect to temperature. The analysis provides the idea of pyrolysis zone for further pyrolysis
4 optimization experiments. The TGA for PP, RBW and 1:1 blend are compared in Fig. 2.
5 From the obtained data, it is clear that the degradation of PP occurred between 240  C and 450  C and the weight loss of 50% takes place at
6 380  C. Whereas, degradation and 50% weight loss for RBW was 260  Ce560  C, 500  C respectively. As expected the value of degradation and
7 50% weight loss of 1:1 blend was observed to be between 250  C and 510  C and 430  C respectively. Therefore, we found that 400e650  C is
8 the suitable pyrolysis zone for further experiments. These observations were consistent with other reported studies of PP and RBW [6,23].
9 A series of experiments were carried out at different temperatures to study the yield of pyrolysis products within obtained pyrolytic zone.
10 The pyrolysis of PP, RBW, and different ratios of PP and RBW (i.e. 1:1, 1:2, 1:3, 2:1, 3:1) yielded three different products i.e. liquid, gas and
11 solid residue. The liquid yield of pyrolysis products of all ratios of PP: RBW has been shown in Fig. 3A. The conducted experiments were
12 performed in the absence of any catalyst, solvents or pressure. It was observed that as the temperature increased from 400  C to 500  C, the
13 liquid yield of PP (1:0) increased and reached to a maximum value of 75.59% at 500  C. While decrease in liquid yield was also observed from
14 550  C to 650  C. Similarly, liquid yield of RBW (0:1) was observed to increase from 400  C to 600  C, reaching maximum yield of 86.5%.
15 Thereafter, the yield was observed to decrease. The feeding material mixture of PP and RBW in different ratios (PP:RBW) i.e. 1:1, 1:2, 1:3, 2:1
16 and 3:1 were also included in study. As the noticeable fact, higher RBW constituent increases liquid yield at higher temperature. Whereas,
17 higher PP content decreases the liquid fraction in co‒pyrolysis blends. Therefore, maximum obtained liquid yield was 80.5% in 1:3 whereas
18 the minimum was 70.07% in 3:1 blend at 550  C. It was also evident that, the maximum liquid yield of PP series were in order of
19 1:0 > 1:1 > 2:1 > 3:1 whereas RBW series were in order of 0:1 > 1:3 > 1:2 > 1:1 (Table 4).
20 The recovery of higher liquid percentage from PP and other PE in similar trend of pyrolysis product distribution were reported earlier
21 [23,24]. However, reported maximum liquid yield of RH pyrolysis with (64.5%) and without (56%) catalyst is less than maximum liquid yield of
22 all ratios obtained in this study [18,25]. Co‒pyrolysis of PP with saw dust, rice husk, and corn cob also shows similar trend and maximum liquid
23 yield at 550  C. Whereas, the maximum reported yield was also less than present study [13]. The better liquid fraction in higher RBW ratio
24 could be due to better cracking of RBW as compared to PP within the fixed bed reactor. Co‒pyrolysis of biomass or biomass derived product
25 with waste plastic has been reported earlier for significant increase in the liquid yield without phase separation due to occurrence of synergistic
26 effect. Several reaction radicals could be formed during co‒pyrolysis reactions at high temperatures such as initiation, depolymerisation,
27 formation of monomers, hydrogen transfer, isomerization and termination by disproportionation or recombination of radicals [15].
28 It was also observed that the percentage of residue decrease with increase in temperature in all experimented ratios which signifies higher
29 degree of conversion (Fig. 3B). Further, the reduction of solid residue was observed with the increase in RBW ratio might occurs due to strong
30 cracking reactions favoured by higher RBW percentage. Several literature states the yield of solid residual product due to pyrolysis or co‒pyrolysis.
31 The decreased overall reaction time with the rise in temperature was also reported earlier [13]. The decomposition of RBW at higher temperatures
32 therefore decreases reaction time and increase optimum temperature yield in blends. The optimum oil was obtained at a reaction time of 30 min.
33 Pyrolysis experiments were also conducted for recovery of various valuable products and series of experiments were conducted in order
34 to optimize the yield. For example, the positive correlation of pyrolysis temperature with quantitative and qualitative yield of both hydrogen
35 and carbon nanotubes synthesis was observed during catalytic pyrolysis of PP [16]. Similarly, BTX aromatics recovery from pyrolysis of PP
36 and PE was also found to be temperature dependent [24].
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65 Fig. 2. TGA curves of PP, 1:1 and RBW.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 6/15

6 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Fig. 3. A. Liquid yield (wt %) and distribution of PP and RBW. B. Residue yield (wt %) and distribution of PP and RBW.
39
40
41 Table 4
42 Maximum obtained product distribution of all ratios (PP:RBW).
43
Sample Optimum Temperature ( C) Liquid Yield (wt. %) þ standard error Residue Yield (wt. %) ± standard error Reaction Time (min)
44
45 PP (1:0) 500 75.59 ± 5.6 5.1 45
RBW (0:1) 600 86.5 ± 2.7 2.8 30
46 1:1 550 73.18 ± 1.4 4.17 40
47 1:2 550 76.5 ± 2.4 4.43 35
48 1:3 550 80.5 ± 3.1 4.12 30
49 2:1 500 71.5 ± 1.8 5.34 40
3:1 550 70.07 ± 2.1 5.2 40
50
51
52
53 3.3. Characterization of liquid pyrolytic yield
54
55 Co‒pyrolysis aims to improve the properties of final blend liquid with respect to individual liquid obtained by raw materials [14]. Table 5
56 clearly shows the significant synergistic effects on physical properties of blend liquid (1:3) in comparison to PP pyrolytic oil in presence of
57 kaolin catalyst, rice bran oil, gasoline and diesel [23, 26e28]. On basis of product distribution experiment series the 1:3 blend was found to
58 be more suitable to carry out further analytical studies and the fuel properties of the pyrolytic oil were analyzed by IS-1448 methods. The
59 density, flash point and fire point of 1:3 blend has been improved while comparing to PP and RBO. Specific gravity and GCV of 1:3 blend was
60 observe to be better than RBO. However, viscosity and carbon residue percentages of 1:3 blend was not found suitable and could be modified
61 in future research work. Most of the fuel properties of 1:3 blend falls within the range of gasoline and diesel properties. Therefore, sub-
62 stantial improvement in fuel properties of blend with respect to PP and RBO is clearly visible and found to be suitable to use as fuel.
63 The overall net energy conversion mainly depends upon carbon and hydrogen percentage of the fuel. As per obtained data the C (wt%), H
64 (wt%) of 1:3 blend was 83.47%, 16.14% respectively which is close to diesel data. However, elemental analysis also shows that the carbon content
65 of 1:3 blend is higher than PP and RBO. On the other hand, the hydrogen percentage of 1:3 blend was comparatively higher than RBO and diesel

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 7/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 7

1 Table 5
2 Comparative physical properties of PP, RBO, 1:3 pyrolytic oil, gasoline and diesel.

3 Properties PP [23] RBO [26] 1:3 Oil Gasoline [27] Diesel [27]
4 Density (Kg/m3) 777.1 922 800.8 720 850
5 Specific Gravity 0.7777 0.920 0.8008 0.72e0.78 0.82e0.85
6 Kinematic Viscosity @100  C (centistokes) e 9.21 0.76 NA N/A
7 Kinematic Viscosity @50  C (centistokes) 2.27 [Cst@ 30  C] 43.52 [Cst@ 40  C] 5.032 N/A 2e5.5 [@40  C]
Conradson Carbon Residue (%) e 0.6 1.03 N/A 0.30
8
Flash Point  C < - 12  C 316 40.75 43 53e80
9 Fire Point  C < - 12  C 337 43 20 to 25 40 to 50
10 Pour Point C < - 45  C 1 22.7 40 40 to 1
11 GCV (MJ/Kg) 47.14 41.1 43.76 42e46 42e45
12 Elemental analysis (% w/w) Present Study [28]

13 C 80.43 76.56 83.47 84.2 86.2


14 H 17.56 15.78 16.14 15.8 13.8
N 0.01 0.02 0.01 0 0
15
S 1.43 0.84 0.32 0 0
16 O 0.57 6.8 0.055 0 0
17 H/C 2.6 2.46 2.3 2.3 1.9
18 O/C 0.0054 0.0667 0.00049 0 0
19
20
21
22 whereas it was lower than PP. However, both the nitrogen and sulfur of 1:3 blend were lower and therefore in safety range. The oxygen content
23 was calculated by mass balance equation. The low oxygen content in 1:3 blend could also address the reduction of corrosion issues [1].
24 Understanding the fuel properties are important as consumer awareness towards biofuel is particularly crucial. The success of green
25 technology alternatives also depends on public adaptation that too when it is being developed worldwide. Density, specific gravity and
26 viscosity of fuel is related with mass properties. These are influential during blending of fuels, calculation of pumping, pressure, injecting of
27 fuel and fuel volume. Similarly, the conradson carbon residue percentage is related to carbon residue after fuel vaporization. This property is
28 useful during engine designing. Flash point is the minimum temperature at which fuel will ignite and fire point is the minimum temperature
29 at which the fuel will remain ignite after removal of source of ignition. These properties indicates the safety during storage and handling
30 with fire hazards. Further, pour point is the minimum temperature at which the fuel could be pumped as the fuel oil constituents could be
31 precipitate as wax during temperature changes. Thereafter, the calorific value is the quantity of heat produced during complete combustion
32 of fuel which indicates the fuel efficiency [2,22].
33 Cold flow properties are another important fuel properties which corresponds to crystallization behaviour in different temperature
34 range. The properties were generally evaluated by the cloud point, cold filter plugging point, and pour point. However, various ASTM
35 methods could be applied for determination of these points but DSC is able to capture crystallization more accurately and less time
36 consuming method. Therefore, the DSC method was applied in this work for cold flow property characterization.
37 Fig. 4 presents the DSC curves of 1:3 blend pyrolytic oil. Two clear peaks and one peak during heating and cooling was observed. The first
38 curve of temperatures at 24  C corresponds to low‒temperature freezing crystals which appears mainly due to presence of methyl esters of
39 unsaturated fatty acids. Whereas another curve at 26  C represents the high temperature freezing crystals, which are associated mainly with
40 methyl esters of saturated fatty acids [20]. However, according to Shadangi et al., 2014, the heating region also corresponds to crystallization
41 behaviour and could be divided in the higher temperature region. It was in between 0 and 30  C and the lower temperature region was in
42 between 0 and ‒30  C. The present graph only shows one peak at 23  C which refers to methyl esters of saturated fatty acids. However,
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65 Fig. 4. DSC curve of 1:3 blend pyrolytic oil under heating and cooling (Heating and cooling rate: 10  C/min).

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 8/15

8 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1 studies report that higher temperature exothermic regions corresponded to the crystallization of the stearin fraction [29]. Literature in DSC
2 crystallization study is very limited but blends of bio-diesel prepared from poultry fat feedstock reports similar finding [20].
3 The chemical characterization of liquid pyrolytic yield was further analyzed by FTIR, GC‒MS and NMR. FTIR is detailed chemical analysis
4 technique to identify various characteristic functional groups by involving infrared ray. According to the fundamental principle, the rays
5 interacts in specific wave length range with molecule in spite of structure, monoatomic elements and homopolar diatomic present in
6 experimental material. The comparative FTIR spectrum for all the three pyrolytic oils (PP oil, RBW oil and 1:3 oil) is shown in Fig. 5 and
7 results found from the transmittance spectrums are presented in Table 6.
8 The C]CH2, CeH stretching vibrations in 3080 cm1 965 cm1, 990 cm1 of the liquid product indicates the presence of vinyl. Whereas
9 CeH stretching vibration on 2850 cm1, 2870 cm1, 2925 cm1 and 2960 cm1 forms Alkyl (methylene) or Alkyl (methyl) in all the liquid
10 products. Similarly, the 1720 cm1 and 1645 cm1 indicates the presence of aldehyde/ketone and acyclic groups respectively. The bands at
11 1450 cm1 with C]C stretching indicates the presence of aromatic group. On the other hand, 723 cm1, 740 cm1, 885 cm1, and 895 cm1
12 with CeH stretching show the presence of aromatic compounds in liquid products. Summarizing, the FTIR spectrum of the 1:3 co-pyrolysis
13 oil closely resemble to RBW rather than PP pyrolysis oil. This could be clearly justified due to presence of three times more RBW amount in
14 1:3 blend. The data of liquid showed the presence of mainly alkyl and aromatic compounds. There were also presence of vinyl, acyclic, and
15 aldehyde/ketone groups.
16 The specific compounds were further identified by matching the corresponding GC‒MS peak from chromatogram with NIST library
17 (Fig. 6). The obtained data represents around 80 different compounds from all the three pyrolytic oil with corresponding area % are enlisted
18 in Table 7. Although compounds could be classified on the basis of several sections but here majorly they were considered as alkane, alkene,
19 amines, imine, ketone, aldehydes, alcohols, ester, acids and other heterocyclic compounds. Approximately 40% of compounds from PP, RBW
20 and 1:3 blend pyrolytic oil belongs to hydrocarbon groups i.e. alkane, alkene and alkyne. Whereas more than 30% of compounds from both
21 PP and 1:3 blends constitutes majorly from oxygen containing groups i.e. Ketone, Aldehydes, Alcohols, Ester and Acids. This percentage of
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44 Fig. 5. FTIR Spectra of PP, RBW, and 1:3 pyrolytic oils.
45
46
47 Table 6
48 FTIR functional groups of PP, RBW, and 1:3 pyrolytic oils.
49 Absorption Peak (cm1) Functional Group Types of Vibrations PP RBW 1:1
50
723 Aromatic CeH bond e √ √
51 740 √ √ √
52 885 e √ √
53 895 √ √ √
965 vinyl CeH bond √ √ √
54
990 e √ √
55 1380 Alkyl (methyl) √ √ √
56 1450 Aromatic C¼C bond √ √ √
57 1645 Acyclic (monosub. alkenes) √ √ √
58 1720 aldehyde/ketone (saturated aliph./ C¼O bond e √ √
cyclic 6-membered)
59
2850 Alkyl (methylene) CeH bond √ √ √
60 2870 Alkyl (methyl) √ √ √
61 2925 Alkyl (methylene) √ √ √
62 2960 Alkyl (methyl) √ √ √
3080 vinyl C¼CH2 bond √ e e
63
64
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 9/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 9

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
Fig. 6. GC-MS of (A) PP pyrolytic oil (B) 1:3 pyrolytic oil (C) RBW pyrolytic oil. Q5
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 10/15

10 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1 Table 7
2 GC-MS analysis of PP, RBW and 1:3 pyrolytic oil.

3 Compound Name Formula PP 1:3 RBW


4 Alkane
5 Cyclopropane C3H6 2.76
6 Cyclobutane C4H8 1.92 4.69
7 3-ethyl-3-azabicyclo[2.1.1]hexane C7H13N 2.72 2.62
Heptane C7H16 6.04 3.55 2.24
8
(1R,2s,3S)-1,2,3-Trimethylcyclopentane C8H16 1.37 2.33 3.09
9 Octane C8H18 1.73
10 4-Methyloctane C9H20 0.81 5.06 3.44
11 1-Undecene C11H22 1.02
12 Undecane C11H24 1.42 0.98
Dodecene C12H24 7.68 4.65 3.08
13 Tetradecane C14H30 3.43 3.40 2.54
14 Icosane C20H42 0.59
15 Alkene
16 2-Butene C4H8 0.63
2-Hexene C6H12 1.10 3.84 2.76
17
(Z)-3-Methyl-2-hexene C7H14 5.74 7.21
18 1-Tetradecene C14H28 2.98 1.05 1.76
19 1-Octadecene C18H36 10.36 6.21
20 Alkyne
21 1-Hexadecyne C16H30 0.54 3.87 1.37
5-Icosyne C20H38 0.16
22 Amines
23 3-aminopropyne C3H5N 0.39
24 Allylamine C3H7N 3.07
25 Aziridine, 1-acetyl-2-methyl- C5H9NO 1.16
Cyclopentanamine C5H11N 2.90
26
3,4,4-Trimethyl-4H-1,2,6-oxadiazine 2-oxide C6H10N2O2 0.66
27 4-Piperidinecarboxylic acid hydrazide C6H13N3O 1.35
28 N-[3-Methylaminopropyl]aziridine C6H14N2 0.60 2.67
29 N-(3-Methyl-2-buten-1-yl)acetamide C7H13NO 0.85
30 N-[(3E)-3-Penten-1-yl]acetamide C7H13NO 0.87 4.83 0.86
N-acetylpiperidine C7H13NO 3.83
31 3-Aminoquinuclidine C7H14N2 0.92
32 2,3-Dimethylpiperidine C7H15N 0.92 0.96
33 Azonane C8H17N 2.88
34 N,1,3-Trimethylcyclopentanamine C8H17N 0.31
1-Isopropyl-2,2-dimethylazetidine C8H17N 1.94
35
1,3-Butanediamine, N,N,N0 ,N0 -tetramethyl- C8H20N2 2.88
36 5-{5-[(Diethylamino)methyl]-1,2,4-oxadiazol-3-yl}-1,3,4-oxadiazol-2-amine C9H14N6O2 1.46 1.79
37 3,3,5-Trimethylcyclohexylamine C9H19N 1.69
38 3-Cyclohexylaminopropylamine C9H20N2 0.69
39 1,1-Bis(1-aziridinyl)-N,N,N0 ,N0 -tetramethylmethanediamine C9H20N4 1.19 1.66 0.84
N-[3-Hexylaminopropyl]aziridine C11H24N2 1.42
40 Imine
41 2,2-Dimethylethylenimine C4H7N 1.85
42 Methanimine,C-ethoxycarbonyl,N-ethoxycarbonylmethyl C8H13NO4 3.43
43 (3Z)eN-Methyl-3-heptanimine C8H17N 1.83
N,N-Diethyl-1-methyl-1,3-propanediamine C8H20N2 0.69
44
N-Isopropyl-2-azido-2-ethylbutanimine C9H19N 1.69
45 N-Octylmethanimine C9H19N 4.34 2.77 1.68
46 Ketone
47 Succinimide C4H5NO2 0.53
48 1-(1H-Imidazole-2-yl)ethanone C5H6N2O 1.02
3-Piperidinone C5H9NO 3.62
49
2,5-Cyclohexadienone C6H6O 0.51
50 1,4-Cyclohexanedione C6H8O2 1.38
51 Spiro [2.4]heptan-4-one C7H10O 1.69 1.02
52 2-Hydroxy-4-methylcyclohexanone C7H12O2 1.58
5-Methylcycloheptane-1,4-dione C8H12O2 1.09 1.67 2.98
53
Bicyclo [3.3.1]nonan-2-one C9H14O 4.31 5.36 6.51
54 3-Hydroxy-4-methyl-3-vinylcyclohexanone C9H14O2 1.95
55 3-(2-hydroxyethyl)-2,3-dimethylcyclopentanone C9H16O2 0.13
56 2-Hexylcyclobutanone C10H18O 3.01
57 Aldehydes
Hexanal C6H12O 3.41 2.77
58
(2Z)-3-(Dimethylamino)-3-(isopropylamino)acrylaldehyde C8H16N2O 3.23
59 Non-2-enal C9H16O 1.92
60 E-2-decenal C10H18O 1.87 1.68 1.72
61 Alcohols
4-Cyclopentene-1,3-diol C5H8O2 5.05
62
Ribitol C5H12O5 2.73 0.85
63 Phenol C6H6O 11.45 3.07
64 Maltol C6H6O3 0.82
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 11/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 11

1 Table 7 (continued )
2 Compound Name Formula PP 1:3 RBW
3
1,2-Dimethylcyclohexanol C8H16O 2.55 3.57
4 3,5-dimethyl-2-octanone C10H20O 1.16
5 Dodecan-1-ol C12H26O 0.79
6 1-Tridecanol C13H28O 2.03
7 Ester and Acids
Formic Acid CH2O2 3.64
8
Isopropenyl acetate C5H8O2 1.91 0.94
9 Methyl 1-hexanesulfonate C7H16O3S 1.84 1.87
10 Ethyl (2E)-3-anilino-2-benzoylacrylate C18H17NO3 1.34
11 Other Heterocyclic Compounds
12 1H-Azepine C6H7N 4.24
2-Methoxytetrahydro-2H-pyran C6H12O2 2.15
13 1,1,2-Trimethylsiletane C6H14Si 1.83
14 5,7,7-Trimethyl-6,8-dioxabicyclo [3.2.1]octane C9H16O2 2.98
15 Octyl 2-pentanyl sulfite C13H28O3S 0.48 0.59
16 Dihydroxycoprostane C27H48O2 3.84
17
18
19
this group slightly increased in RBW pyrolytic oil, i.e. ~36%. However, nitrogen containing groups are similar in percentage of PP and 1:3
20
blend pyrolytic oil i.e. (~20%). But it is around 15% in RBW. Moreover, sulphur compounds as impurities were also found in RBW and 1:3
21
blend pyrolytic oil in very small quantity (<2%).
22
Presence of high amount of alkane and alkene in PP pyrolytic oil such as Heptane, Dodecene and 1‒Octadecene were reported in several
23
literature. However, plastic polymers were also known for hydrocarbon donor in blend pyrolysis experiment [10]. Although, no study claims
24
regarding pyrolysis of RBW yet. But studies from other waxes and RH hints the degradation of waxes, cellulose and hemicellulose results into
25
higher hydrocarbons and oxygen containing groups [17]. The investigation of chromatograms showed the carbon distribution of pyrolytic oil
26
mainly ranging from C4 to C16. However, very few lower or higher carbon compounds were also observed. Certainly, this range of compounds
27
were suitable to be used as fuel [17].
28 1
H NMR experiments were carried out to confirm the functional group of pyrolytic oil component. Fig. 7 showed the 1H NMR spectrum of
29
all pyrolytic oil and were further elaborated in Table 8. The comparatively higher peaks were appeared in the range of 0.5 ppme2.5 ppm. The
30
difference in chemical makeup of all pyrolytic oil could be easily notable. The table was entirely based on general region of chemical shift
31
studies by Silverstein et al., 2014 [30]. The chemical shifts of protons in all the three pyrolytic oil compounds fall roughly into aliphatic region
32
whereas aromatics or heteroaromatic compounds constitutes very less in percentages. It is also now known from the studies that aliphatic
33
compounds are flammable, allowing the use of hydrocarbons as fuel [22].
34
Aliphatic alicyclic compounds in PP and RBW appears to be highest in overall constituents whereas aliphatic alicyclic along with b‒
35
substituted aliphatic was higher in 1:3 blend pyrolytic oil. From 0.5 ppm to 2.0 ppm region, the spectra proton of PP, 1:3 blend, RBW
36
constituted around 77%, 68% and 73% respectively. From 2.0 ppm to 3.0 ppm region, the spectra proton were alkynes, a‒monosubstituted
37
aliphatic and a‒disubstituted aliphatic compounds which was higher in 1:3 blend in comparison to both pyrolytic oil. The alkene, alkyne
38
chemical shifts were from 4.5 to 7.5 and 2.0 to 3.0 respectively [30]. The results of 1H NMR is consistent with the early discussion with GC-MS
39
i.e. alkene trends follows 1:3 blend < PP < RBW whereas alkyne follows PP < RBW<1:3. The obtained results were found to be similar with
40
other RBW studies [6].
41
42
43 3.4. Characterization of residual char from blend
44
45 The physical characterization of residual char of 1:3 blend was done in order to know the utility as cheap absorbent, carbon coating, solid
46 fuel, direct fuel, household briquette, carbon sequestration and soil amendment. The proximate and ultimate analysis of char has been
47 compared in Table 3. The proximate analysis includes the determination of moisture, volatile matter, ash and fixed carbon content which
48 were 1.57%, 27.8%, 19.3% and 51.33% respectively. Similarly, ultimate analysis was used to determine the basic elemental composition of
49 carbon, hydrogen, nitrogen, sulfur and oxygen. Obtained char constitutes 92.6% carbon, 1.36% hydrogen, 0.04% nitrogen, 0.83% sulfur and
50 5.17% oxygen. Whereas, GCV was determined to be 27.26 MJ/kg.
51 SEM images of 1:3 blend char taken at 1000 and 2000 magnifications are presented as Fig. 8A,B. The char exhibited a heterogeneous
52 distribution of pores and a rough texture. The average size of pores on the surface was 13.97 mm. Volatile matters escape during pyrolysis of
53 biomass which appears as pores in the surface of the char. But according to 1000x resolution, it could be concluded that the char is not highly
54 porous comparatively. This result is consistent with BET analysis. The surface area of the char sample as per BET equation, was calculated as
55 12.621 m2/g. The obtained char has relatively higher surface area compared to the red oak char and co-pyrolysis residual char from red oak/
56 HDPE i.e. 3.80 and 6.75 m2/g, respectively [31]. But, physical or chemical treatment could increase further porosity and quality.
57 Poor quality RH char due to high silica content in surface was reported to improve by leaching steps [3]. Whereas improved char quality
58 was also reported from co-pyrolysis of red oak biomass with HDPE [31]. However, the RH bio‒char for improving carbon sequestration and
59 soil amendment capabilities were also discussed in literature [32]. Similarly, reports from application of HDPE plastic char as briquette fuel
60 and waste CDs/DVDs char for iron carburization were also suggested [31,33]. Therefore, carbonaceous char could be used by various ways.
61
62 3.5. Practical implications of this study
63
64 Present study contribute towards the exploration of sustainable and renewable biofuel alternative. Depletion of fossil fuel reservoirs,
65 dramatic environmental changes, rapidly growing industrialization and energy demand are few causes of this research enthusiasm.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 12/15

12 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 Fig. 7. The H1 NMR spectra of PP, 1:3 blend and RBW pyrolytic oil.
64
65

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 13/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 13

1 Table 8
2 NMR analysis of polypropylene pyrolytic oil.

3 Functional groups Chemical shift (ppm) Percentage of total hydrogen (%)


4 PP 1:3 RBW
5
a-disubstituted aliphatic, Alkene, Aromatic and heteroaromatic 9.0e6.0 2.8 3.32 3.38
6
a-disubstituted aliphatic, Alkene 6.0e5.0 2.05 6.88 3.48
7 a-monosubstituted aliphatic, a-disubstituted aliphatic, Alkene 5.0e4.0 7.94 1.83 8.54
8 Alkynes, a-monosubstituted aliphatic, a-disubstituted aliphatic 3.0e2.0 9.45 20.03 11.38
9 Aliphatic alicyclic, b-substituted aliphatic 2.0e1.0 35.7 44.25 22.56
Aliphatic alicyclic 1.0e0.5 42.08 23.69 50.66
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65 Fig. 8. A. SEM of 1:3 char at 2000X. B. SEM of 1:3 char at 1000X.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 14/15

14 Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15

1 Several reports on the innovative biofuel resources has been reported yet but practically fails the criteria of ‘sustainability’. Therefore, the
2 present work involved the pyrolytic oil production from RBW and PP. According to CFTRI, Mysore, several tonnes of unpurified RBW has no
3 market value and thus are counted in industrial loss. On the other hand, purification of wax involves several chemicals which are not only
4 included with industrial water effluent but the final product are also not safe to be used as food additives [34]. Pesticides and heavy metal
5 contamination has also been reported from rice bran extract [5]. Similarly, the over use of plastic material has raise the concern of envi-
6 ronmental pollution and recycling of waste plastic into valuable asset. PP alone adds approximately 24.3% plastic excess in municipal solid
7 waste [24].
8 Depending upon biomass feedstocks, the pyrolytic oil contains high oxygen percentage causes low GCV, corrosion or stability issues.
9 Possible upgrading process could be adding catalyst, solvent or hydrogen sources. These process could not be much economical [1]. Hence,
10 the present study also aims to discuss the upgradation of the obtained pyrolytic liquid i.e. improvement of liquid characteristics by co-
11 pyrolysis method. Thereafter, the pyrolytic liquid have advantage of easy storage and transportation benefits. The pyrolytic liquid could
12 be used as precursors of valuable chemical extraction, used to run engines, turbines, boilers in industries and could also blend with fossil fuel
13 for tremendous use.
14
15 4. Conclusion
16
17 Summarizing, 1:3 blend pyrolytic oil at 550  C of PP and RBW was found most suitable blend having optimum yield of 80.5% with calorific
18 value of 43.76 MJ/kg. Analytical results from FTIR, GC-MS, NMR supports the higher aliphatic compounds in liquid yield. In addition, the
19 physical properties of the pyrolytic oil were fairly comparable to Diesel and Gasoline. Unfortunately, the obtained char from 1:3 blend
20 residue was not much porous. But further simple physical or chemical treatment studies could improve the quality of char. Therefore, the
21 study could solve fossil fuel dependency, improving waste management of plastic, upgradation of pyrolysis oil and hence are overall
22 economic.
23
24
Acknowledgement
25
26
The authors are grateful for the financial research support to Department of Chemical Engineering, National Institute of Technology,
27 Q3
Rourkela, India.
28
29
30 References
31
[1] Huiyan Zhang, Jianlong Nie, Rui Xiao, Baosheng Jin, Changqing Dong, Guomin Xiao, Catalytic co-pyrolysis of biomass and different plastics (polyethylene, polypropylene,
32 and polystyrene) to improve hydrocarbon yield in a fluidized-bed reactor, Energy Fuels 28 (3) (2014) 1940e1947.
33 [2] Namrata Kumari, Raghubansh Kumar Singh, Biodiesel production from local mixed algal culture of Rourkela, Odisha, J. Biochem. Technol. 7 (1) (2016) 1078e1083.
34 [3] Cristina Deiana, Dolly Granados, Venturini Rosa, Alejandro Amaya, Marta Sergio, Ne stor Tancredi, Activated carbons obtained from rice husk: influence of leaching on
textural parameters, Ind. Eng. Chem. Res. 47 (14) (2008) 4754e4757.
35 [4] Siti Zullaikah, Chao-Chin Lai, Shaik Ramjan Vali, Yi-Hsu Ju, A two-step acid-catalyzed process for the production of biodiesel from rice bran oil, Bioresour. Technol. 96
36 (17) (2005) 1889e1896.
37 [5] Amended final report on the safety assessment of oryza sativa (rice) bran oil, oryza sativa (rice) germ oil, rice bran Acid,Oryza sativa (rice) bran wax, hydrogenated rice
bran wax, oryza sativa (rice)Bran extract, oryza sativa (rice) extract, oryza sativa (rice) germ powder, oryza sativa (rice) starch, oryza sativa (rice) bran, hydrolyzed rice
38 bran extract, hydrolyzed rice bran protein, hydrolyzed rice extract, and hydrolyzed rice protein, Int. J. Toxicol. 25 (Suppl. 2) (2006) 91e120. https://doi.org/10.1080/
39 10915810600964626 PMid:17090480.
40 [6] Qing Liu, Peng Liu, Zhi-Xiang Xu, He Zhi-Xia, Qian Wang, Bio-fuel oil characteristic of rice bran wax pyrolysis, Renew. Energy 119 (2018) 193e202.
[7] Arndt Asperger, Engewald Werner, Gerd Fabian, Analytical characterization of natural waxes employing pyrolysisegas chromatographyemass spectrometry, J. Anal.
41
Appl. Pyrol. 50 (2) (1999) 103e115.
42 [8] Chi Diem Doan, Chak Ming To, Mike De Vrieze, Frederic Lynen, Sabine Danthine, Allison Brown, Koen Dewettinck, R. Ashok, Patel, Chemical profiling of the major
43 components in natural waxes to elucidate their role in liquid oil structuring, Food Chem. 214 (2017) 717e725.
44 [9] Yogesh Chandrakant Rotliwala, Parimal Amaratlal Parikh, Thermal degradation of rice-bran with high density polyethylene: a kinetic study, Kor. J. Chem. Eng. 28 (3)
(2011) 788e792.
45 [10] Jayeeta Chattopadhyay, T.S. Pathak, R. Srivastava, A.C. Singh, Catalytic co-pyrolysis of paper biomass and plastic mixtures (HDPE (high density polyethylene), PP
46 (polypropylene) and PET (polyethylene terephthalate)) and product analysis, Energy 103 (2016) 513e521.
47 [11] Piotr Rutkowski, Chemical composition of bio-oil produced by co-pyrolysis of biopolymer/polypropylene mixtures with K2CO3 and ZnCl2 addition, J. Anal. Appl. Pyrol.
95 (2012) 38e47.
48 [12] Dadi V. Suriapparao, Deepak Kumar Ojha, Tanumoy Ray, R. Vinu, Kinetic analysis of co-pyrolysis of cellulose and polypropylene, J. Therm. Anal. Calorim. 117 (3) (2014)
49 1441e1451.
50 [13] J.L. Ye, Q. Cao, Y.S. Zhao, Co-pyrolysis of polypropylene and biomass, Energy Sources, Part A 30 (18) (2008) 1689e1697.
[14] Levent Ballice, Classification of volatile products evolved from the temperature-programmed co-pyrolysis of Turkish oil shales with atactic polypropylene (APP), Energy
51 Fuel. 15 (3) (2001) 659e665.
52 [15] Levent Ballice, Rainer Reimert, Temperature-programmed co-pyrolysis of Turkish lignite with polypropylene, J. Anal. Appl. Pyrol. 65 (2) (2002) 207e219.
53 [16] Jie Liu, Zhiwei Jiang, Haiou Yu, Tao Tang, Catalytic pyrolysis of polypropylene to synthesize carbon nanotubes and hydrogen through a two-stage process, Polym. Degrad.
Stabil. 96 (10) (2011) 1711e1719.
54 [17] Yao Lu, Xian-Yong Wei, Jing-Pei Cao, Li Peng, Fang-Jing Liu, Yun-Peng Zhao, Xing Fan, et al., Characterization of a bio-oil from pyrolysis of rice husk by detailed
55 compositional analysis and structural investigation of lignin, Bioresour. Technol. 116 (2012) 114e119.
56 [18] Leiyu Zhou, Hongmin Yang, Hao Wu, Meng Wang, Daqian Cheng, Catalytic pyrolysis of rice husk by mixing with zinc oxide: characterization of bio-oil and its rheological
behavior, Fuel Process. Technol. 106 (2013) 385e391.
57
[19] Ji-lu ZHENG, Pyrolysis of rice husk and sawdust for liquid fuel, J. Environ. Sci. 18 (2) (2006) 392e396.
58 [20] Manuel Garcia-Perez, Thomas T. Adams, John W. Goodrum, K.C. Das, Daniel P. Geller, DSC studies to evaluate the impact of bio-oil on cold flow properties and oxidation
59 stability of bio-diesel, Bioresour. Technol. 101 (15) (2010) 6219e6224.
60 [21] N. Othman, N.E.A. Basri, M.N.M. Yunus, L.M. Sidek, Determination of physical and chemical characteristics of electronic plastic waste (Ep-Waste) resin using proximate
and ultimate analysis method, in: International Conference on Construction and Building Technology, 2008, pp. 169e180.
61 [22] Tanmya Rout, Debalaxmi Pradhan, R.K. Singh, Namrata Kumari, Exhaustive study of products obtained from coconut shell pyrolysis, J Environ Chem Eng 4 (3) (2016)
62 3696e3705.
63 [23] Achyut K. Panda, R.K. Singh, Experimental optimization of process for the thermo-catalytic degradation of waste polypropylene to liquid fuel, Adv Energy Eng 1 (3)
(2013) 74e84.
64 [24] Su-Hwa Jung, Min-Hwan Cho, Bo-Sung Kang, Joo-Sik Kim, Pyrolysis of a fraction of waste polypropylene and polyethylene for the recovery of BTX aromatics using a
65 fluidized bed reactor, Fuel Process. Technol. 91 (3) (2010) 277e284.

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011
JOEI479_proof ■ 13 August 2018 ■ 15/15

Akancha et al. / Journal of the Energy Institute xxx (2018) 1e15 15

1 [25] Y. Lu, X.-Y. Wei, F.-J. Liu, Z.-M. Zong, L.-C. Rong, Y.-P. Zhao, X. Fan, et al., Evaluation of an upgraded bio-oil from the pyrolysis of rice husk by acidic resin-catalyzed
esterification, Energy Sources, Part A Recovery, Util. Environ. Eff. 36 (6) (2014) 575e581.
2
[26] Shailendra Sinha, Avinash Kumar Agarwal, Sanjeev Garg, Biodiesel development from rice bran oil: transesterification process optimization and fuel characterization,
3 Energy Convers. Manag. 49 (5) (2008) 1248e1257.
4 [27] Sachin Kumar, Conversion of Waste High-density Polyethylene into Liquid Fuels, PhD diss., 2014.
5 [28] R. Hilten, R. Speir, J. Kastner, K.C. Das, Production of aromatic green gasoline additives via catalytic pyrolysis of acidulated peanut oil soap stock, Bioresour. Technol. 102
(17) (2011) 8288e8294.
6 [29] Krushna Prasad Shadangi, Kaustubha Mohanty, Production and characterization of pyrolytic oil by catalytic pyrolysis of Niger seed, Fuel 126 (2014) 109e115.
7 [30] Robert M. Silverstein, Francis X. Webster, David J. Kiemle, David L. Bryce, Spectrometric Identification of Organic Compounds, John wiley & sons, 2014.
8 [31] Yuan Xue, Shuai Zhou, Robert C. Brown, Atul Kelkar, Xianglan Bai, Fast pyrolysis of biomass and waste plastic in a fluidized bed reactor, Fuel 156 (2015) 40e46.
[32] Salman Raza Naqvi, Yoshimitsu Uemura, Noridah Osman, Suzana Yusup, Production and evaluation of physicochemical characteristics of paddy husk bio-char for its C
9 Sequestration applications, BioEnergy Res 8 (4) (2015) 1800e1809.
10 [33] Irshad Ahmed Mansuri, Rita Khanna, Ravindra Rajarao, Veena Sahajwalla, Recycling waste CDs as a carbon resource: dissolution of carbon into molten iron at 1550 C,
11 ISIJ Int. 53 (12) (2013) 2259e2265.
[34] Purified Wax from Rice Bran Wax Sludge, CENTRAL FOOD TECHNOLOGICAL RESEARCH INSTITUTE MYSORE-570 020, INDIA, modified at 2016-07-22, retrieved at 2018-
12 07-21, from http://www.cftri.com/technologies/CONP/pwr.pdf.
13
14

Please cite this article in press as: Akancha, et al., Co-pyrolysis of waste polypropylene and rice bran wax‒ production of biofuel and its
characterization, Journal of the Energy Institute (2018), https://doi.org/10.1016/j.joei.2018.07.011

You might also like