Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Article

Cite This: Energy Fuels 2020, 34, 460−473 pubs.acs.org/EF

Catalytic Co-pyrolysis of Biomass and Plastics (Polypropylene and


Polystyrene) Using Spent FCC Catalyst
Kanduri Praveen Kumar and Seethamraju Srinivas*
Department of Energy Science and Engineering, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
*
S Supporting Information

ABSTRACT: In situ catalytic co-pyrolysis (CCP) of mixtures of lignocellulosic biomass (LB) and plastic (polypropylene (PP)
and polystyrene (PS)) with spent FCC catalyst (10% w/w loading) was performed in a semibatch reactor. The experiments
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

were carried out using 30 g of feed at 510 °C and three different feed ratios of LB to plastic (1:2, 1:1, and 2:1). The effect of
Downloaded via INST OF CHEMICAL TECHNOLOGY on March 14, 2024 at 07:56:07 (UTC).

feed ratio on bio-oil yield, composition, and carbon yield of products (mainly, aromatics and olefins) was evaluated. A maximum
aromatic selectivity of ∼49%C with PP and ∼82%C with PS was observed at a feed ratio of 0.5. Co-processing LB with PP
produced more valuable chemicals like toluene and xylene at equal mixing ratios. The total yield of monoaromatics and olefins
was maximum (∼73%C) at an LB to PP ratio of 0.5. When LB was co-fed with PS, the more valuable styrene monomers and
ethylbenzene were found significantly in the product at intermediate mixing ratios. While co-feeding PP with LB improved the
(H/C)eff ratio of the pyrolysis liquid, the change in the (H/C)eff ratio was insignificant in the case of biomass and PS mixtures.
The spent FCC catalyst helps in oxygen removal from the bio-oil and improves its quality (higher aromatic fraction) compared
to oil obtained from thermal co-pyrolysis of the biomass and plastic mixtures.

1. INTRODUCTION unstable. Also, oxygenates initiate polymerization reactions and


The global primary energy demand grew by 2.9% in 2018 and promote aging of bio-oil upon storage. Therefore, upgrading of
was driven primarily by fossil fuels. On average, the global oil bio-oil is necessary for its acceptance as a replacement for
consumption increased by 1.4 million barrels per day (1.5%), petroleum fuels.2,11,12 Over the past few decades, research has
which led to a growth in carbon dioxide emissions by 2.0%.1 focused on the use of catalysts in biomass pyrolysis to produce
Therefore, in comparison to fossil resources, lignocellulosic deoxygenated pyrolysis liquid. However, poor hydrocarbon
biomass (LB) can be used as an alternative on account of its yields and high char formation are the major drawbacks of this
renewability, abundant availability, and CO2 neutrality.2,3 Over process.13−15 Since biomass is hydrogen-deficient (low (H/
the past decade, LB has contributed significantly to the C)eff ratio of ∼0.3 only), it results in a low hydrocarbon yield.
renewable energy sector by meeting approximately 10% of the Further, most of the hydrogen results in water formation
global energy demand.4 during the catalytic deoxygenation process. The H2-deficient
Among different biomass conversion alternatives like environment also leads to rapid char formation, which
pyrolysis, gasification, and hydrothermal liquefaction, pyrolysis eventually deactivates the catalyst.16,17
offers an attractive way of transforming the low-energy density Since plastics are predominantly hydrocarbons and H2-rich,
biomass into a high-energy-density liquid called bio-oil. The co-feeding plastics with biomass during pyrolysis is envisaged
thermal breakdown of polymeric biomass components to overcome H2 deficiency. Also, the bio-oil from this process
(cellulose, hemicellulose, and lignin) into smaller fragments has better stability and heating value compared to oil produced
occurs at moderate to relatively high temperatures in an inert from the pyrolysis of LB alone.18,19 Most plastics are
atmosphere during biomass pyrolysis.5−7 The pyrolysis predominantly non-oxygenated hydrocarbons and can serve
products include bio-oil, bio-char, and gases like hydrogen, as low-cost hydrogen sources during co-pyrolysis with LB. Up
carbon monoxide, methane, and carbon dioxide. Owing to the to 2015, approximately 6300 MMT of plastic waste was
heterogeneity of biomass, the bio-oil yield and composition generated; 79% of it was landfilled, 12% incinerated, and only
varies with the type of feedstock and process conditions 9% recycled.20 One of the advantages of the co-pyrolysis
(temperature, heating rate, and residence time). Research process is the utilization of waste plastics (including end-of-
suggests that bio-oil typically contains 45−50 wt % oxygen and life-plastics) and waste biomass as feedstocks; this can reduce
is distributed in oxygenates like hydroxyaldehydes, hydrox- the plastic waste sent to landfill sites. Co-pyrolysis, thus, can
yketones, carboxylic acids, alcohols, and phenolics. The also be an effective waste management strategy.
presence of oxygen in the bio-oil makes it behave differently Conversion of biomass to high-value products needed in the
from hydrocarbon fuels (e.g., heavy oil), which contain less petrochemical industry instead of fuels is the focus of most
than 1.0 wt % oxygen.2,8 Due to the presence of oxygenates, research today; these include olefins and aromatics like
the bio-oil has a lower heating value (16−19 MJ/kg) compared
to conventional fuels such as diesel or gasoline (40−45 MJ/ Received: September 13, 2019
kg).9,10 Further, oxygenates in the bio-oil impart high acidity, Revised: December 10, 2019
corrosiveness, and high viscosity, which collectively make it Published: December 10, 2019

© 2019 American Chemical Society 460 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

Figure 1. Schematic diagram of the co-pyrolysis experimental setup.

benzene, toluene, and xylene, which are the starting materials proportion on the %C selectivity in the bio-oil, mainly,
for manufacturing a wide range of specialty chemicals.9,17 aromatics, olefins, and oxygenates, was investigated.
Catalytic co-pyrolysis (CCP) of biomass and plastics has an
advantage over thermal co-pyrolysis as the reaction mechanism 2. MATERIALS AND METHODS
can be tailored to produce specific fuels and chemicals since 2.1. Raw Materials. Groundnut shells were procured from the
several reactions occur at the catalyst sites, for example, local market area, finely ground to a particle size of ∼1 mm, and dried
dehydration, aromatization, isomerization, oligomerization, for 24 h in an oven at 105 ± 2 °C to remove the moisture.
alkylation, and cyclization.21 Various zeolite catalysts, and in Commercially available polypropylene (PP) and polystyrene (PS)
particular, HZSM-5, have been extensively used in the beads (∼4 mm in diameter) were also purchased from the local
pyrolysis studies for the production of petrochemicals.22−33 market area. The catalyst used was spent FCC catalyst with a Si/Al
ratio of 1.29 and was supplied in kind by L&T India. The catalyst was
Some of these studies are summarized in Table S1 in the used directly in the experiments without any pretreatment.
Supporting Information. Spent FCC catalyst is also a zeolite 2.2. Experimental Setup and Method. Figure 1 shows the
that can be used in biomass co-pyrolysis. FCC catalyst is used schematic of the experimental setup used for the catalytic co-pyrolysis
in the petroleum refineries to convert heavy petroleum (CCP). The reactor is a 15 cm long quartz tube, with 5 cm inner
fractions into lighter products such as gasoline, diesel, and diameter. An electrical tubular furnace was used to heat the reactor to
other products. Globally, 150,000−170,000 tons of spent FCC the desired temperature, and the set point temperature was
maintained using a PID controller. A condenser with a cooling
catalyst is generated as a waste per year and is typically water jacket was connected at the exit of the reactor to cool the
disposed of in landfills.34 The reuse of this spent FCC catalyst pyrolysis vapors. The condensed liquid (bio-oil) was collected in a
is expected to decrease the costs involved in the procurement round-bottom flask, and the noncondensable gases in gas bags for
of fresh catalyst, reduce efforts in the development of analysis.
proprietary or new catalysts, and improve overall process Initially, 15−30 g of feed mixture consisting of finely ground LB
economy. and plastic was thoroughly mixed with the catalyst (10% w/w of feed)
and loaded into the cylindrical quartz reactor. Subsequently, the
Groundnut is one of the world’s leading oilseed crops with
reactor was placed in the electric tubular furnace and purged using N2
an estimated production of ∼46.58 MMT in 2017−2018 to ensure the presence of an inert atmosphere before starting the
according to the USDA.35 Groundnut shells (GS) are a low- experiment. The reactor temperature was set to 510 °C at a heating
cost byproduct of groundnut processing and are a potential rate of 10 °C/min and a holding time of 1 h. The operating
source of waste biomass for producing renewable fuels. In the temperature of 510 °C was selected as the major components of
global context, India is the second-largest producer of lignocellulosic biomass (cellulose, hemicellulose, and lignin) are
groundnuts with an estimated production of ∼5.2 MMT in completely pyrolyzed at this temperature.2 Also, this is close to the
temperature used or is the average value of the temperature range
2018.36 In India, groundnuts are grown throughout the year reported in most studies in Table S1. After the completion of the
and have the ability to provide residual biomass throughout the process, the reactor was allowed to cool down to room temperature,
year; this alleviates the seasonal biomass availability issue, and the weights of liquid and char were measured. The percentage
which is otherwise one of the hurdles in biomass utilization. yields of liquid, char, and gases were calculated as given in eqs 1−3.
This study aims to investigate the yield of aromatics from Likewise, each experiment was performed in triplicate to ensure
catalytic co-pyrolysis of groundnut shell (GS) and commonly repeatability and reproducibility.
2.3. Product Analysis. The pyrolysis oil was analyzed using an
used plastics (polypropylene (PP) and polystyrene (PS)) using Agilent 7890 gas chromatograph (GC) equipped with a Jeol
spent FCC catalyst. The novelty of the work lies in trying to AccuTOF GCV mass spectrometer (MS). An HP-5 capillary column
valorize waste biomass and plastic streams using a waste (30 m × 0.25 mm ID × 0.25 μm film thickness) was used to separate
refinery stream (spent FCC catalyst). The effect of plastic the products using He as a carrier gas (flow rate of 1.0 mL/min). The

461 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

split ratio was 10:1. The GC inlet and the MS ion source temperature Feedstocks with a high EHI [(H/C)eff > 1] are easier to
was 250 °C. The electron impact ionization was 70 eV. The mass convert to hydrocarbons, whereas feedstocks with a lower EHI
detection range of MS was 35−650 amu. The GC oven was [(H/C)eff < 1] result in poor hydrocarbon yields and cause
programmed to reach 50 °C in the first 2 min, then heated to 140 °C rapid catalyst deactivation due to coke formation during
at 5 °C/min and held for 3 min, further heated to 240 °C at 5 °C/min
and held for 3 min, and finally heated to 280 °C at 10 °C/min. For
CCP.16,19 PP has a much higher (H/C)eff ratio of 2.1
each analysis, 0.4 μL of the liquid sample was injected into the GC. compared to those of LB and PS, which have (H/C)eff values
The pyrolysis gas samples were analyzed offline using a gas of 0.19 and 0.82, respectively. These values reveal that plastics,
chromatograph (NUCON-5765, India) equipped with a thermal especially PP, are H2-rich, while biomass is H2-deficient.
conductivity detector. A Carbosieve column was used to separate H2,
H − 2O − 3N − 2S
N2, CH4, CO, and CO2. Argon was used as the carrier gas, and a EHI or (H/C)eff =
standard gas mixture with known concentrations of H2, N2, CH4, CO, C (7)
and CO2 was used as the reference to find the composition of where C, H, O, N, and S indicate the number of atoms per unit
pyrolysis gases. For each analysis, 0.5 μL of gas sample was injected
into the GC. weight of carbon, hydrogen, oxygen, nitrogen, and sulfur,
respectively. The proximate analysis of the LB, PP, and PS
3. RESULTS AND DISCUSSION samples is given in Table S2 of the Supporting Information.
3.2. Thermogravimetric Analysis (TGA). The thermal
The following calculations were used in this study: decomposition behavior of LB, PP, and PS was evaluated by a
weight of liquid thermogravimetric analyzer (model no. STA 449 F3 Jupiter
liquid yield (%) = × 100 NETZSCH). For each run, approximately 15 mg of the sample
total weight of the feed (1)
was placed in an alumina crucible and pyrolyzed under a N2
weight of Char atmosphere (80 mL/min flow rate) at a heating rate of 10 °C/
char yield (%) = × 100 min from 35 to 600 °C for plastic samples and 35 to 950 °C
total weight of the feed (2)
for the LB sample. For the LB sample, the N2 atmosphere was
gas yield (%) = 100 − %liquid yield − %char yield (3)
switched to the O2 atmosphere at 800 °C to determine the
amount of fixed carbon. It was observed that LB decomposed
carbon yield (%C) at a lower temperature than plastics; the degradation
temperature range of LB was ∼200−400 °C, and the
number of carbon atoms in a functional group
= corresponding values are ∼350 to 480 °C and ∼350 to 450
total number of carbon atoms in all the functional groups °C for PP and PS, respectively (see Figure S1 in the
× 100 (4) Supporting Information). The decomposition temperature
range for PP was higher compared to PS.
aromatic selectivity (%C) 3.3. Thermal Pyrolysis of Individual and Mixed Feeds
number of carbon atoms in an aromatic product of LB-PP and LB-PS. Figure 2 summarizes the product yields
= from thermal pyrolysis of individual LB and PP samples and
total number of carbon atoms in all the aromatic products
their mixtures at three different mixing ratios.
(5)
× 100 Thermal pyrolysis (TP) of PP produced ∼62 wt % liquid
yield and ∼38 wt % gases. There was no char formation
olefin selectivity (%C)
observed in the pyrolysis of PP. The primary pyrolysis
number of carbon atoms in an olefin product products of PP were mainly branched olefins and alkanes,
=
total number of carbon atoms in all the olefin products which result from the random scission of the parent polymer
× 100 (6)
Equation 4 represents the %C for each functional group like
olefins, aromatics, alkanes, etc. Equations 5 and 6 represent the
selectivity of each component within the same functional
group, for example, benzene, toluene, etc., in aromatics.
3.1. Characteristics of Raw Materials. The elemental
analysis of both the biomass and polymer beads was carried
out using T withhermo Finnigan, Italy Flash EA 1112 series
analyzer and is shown in Table 1. The effective hydrogen index
(EHI) or (H/C)eff is an indicator of net hydrogen to carbon
ratio in a feedstock and is calculated according to eq 7.37

Table 1. Elemental Composition of Biomass, Polypropylene


(PP), and Polystyrene (PS) Used

element biomass (wt %) PP (wt %) PS (wt %)


C 47.89 85.19 93.63
H 6.57 14.81 6.36
N 1.13
S
O 44.42 Figure 2. Product yields as a function of feed ratio for thermal
(H/C)eff 0.19 2.1 0.82 pyrolysis (LB and PP).

462 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

chain. The olefins accounted for approximately 76% of the content with the change in the feed ratio are not so significant,
total carbon yield (see Figure 3). In the case of thermal which indicates that there are minimal interactions between LB
and PP components for aromatic formation in the absence of
the catalyst. Therefore, TCP has a minimal effect in improving
the aromatic distribution in the bio-oil.
The thermal pyrolysis (TP) of PS predominantly produced a
liquid product rich in aromatics. The liquid, gas, and char
yields are ∼88, ∼4, and ∼8 wt %, respectively (see Figure 4).
Styrene monomer was the most abundant component found in
the liquid from thermal pyrolysis of PS followed by methyl
styrene, toluene, and ethylbenzene.

Figure 3. Distribution of functional groups as a function of feed ratio


for thermal pyrolysis (LB and PP).

pyrolysis of LB, the bio-oil yield was ∼35 wt %, and the gas
yield was ∼28 wt %. Besides, there was significant char
formation (∼37 wt %), which is higher than the liquid and gas
yields. Thermal pyrolysis of LB produced predominantly
oxygenated compounds: carboxylic acids, ketones, furans,
aldehydes, and phenols. These products are mostly the
degraded fragments of cellulose, hemicellulose, and lignin.
Figure 4. Product yields as a function of feed ratio for thermal
Thermal co-pyrolysis of LB and plastic mixtures (TCP)
pyrolysis (LB and PS).
produced different compounds when compared to pyrolysis of
the individual components only. For TCP of LB and PP
mixtures, the liquid yield was ∼50 wt % at an LB/PP ratio of TCP of LB and PS mixtures produced a higher liquid yield
0.5 and ∼32 wt % at a ratio of 2 (see Table S3 in the compared to TCP of LB-PP mixtures. The liquid yield values
Supporting Information for phase distribution values between in TCP of LB and PS mixtures were in the range of 45−64 wt
the organic and the aqueous layers). This is in line with the % with the maximum yield observed at a high PS loading (LB/
expectation since higher amounts of plastic in the mixture PS ratio of 0.5) (see Table S3 in the Supporting Information
result in higher liquid yields. The gas yields for TCP for the phase distribution values). The char yield was
experiments were relatively higher than the TP of the proportional to the biomass fraction in the feed mixture, that
individual components, which indicates the presence of is, higher char yields at a high LB loading (see Figure 4).
significant interactions between LB and PP during co-pyrolysis. Figure 5 summarizes the product distribution in TCP of LB
As seen in Figure 3, which shows the effect of feed ratio on and PS mixtures. The aromatic yields almost remained
the distribution of pyrolysis products in TCP of LB and PP constant and did not change appreciably with changes in the
mixtures, the aromatic content increased with the increase in feed ratio. The aromatics yield was in the range of 77−80%C.
LB fraction in the feed. The aromatic yield (∼35%C) was Similar to PS pyrolysis, styrene monomer, methyl styrene,
maximum at an LB/PP ratio of 2 and minimum (∼24%C) at a toluene, and ethylbenzene were found abundantly in the bio-
ratio of 0.5. However, the aromatics are predominantly oil. The selectivity of phenols (product of lignin degradation
phenolic compounds and are observed at all the feed ratios. from biomass) did not change with the feed ratio; these results
Most of the phenolic compounds are the degradation products suggest that the change in the feed ratio did not affect the
of lignin, for example, catechol, methyl catechol, and product distribution in the bio-oil to a significant extent in
resorcinol. The nonoxygenated monoaromatic (NOMA) TCP. The advantage of using a catalyst in the co-pyrolysis
selectivity increased with LB fraction in the feed, the maximum process is discussed in the next few sections.
and minimum selectivity values being ∼22%C and ∼7%C at 3.4. Effect of Feed Composition on Yields and
the feed ratios of 2 and 0.5, respectively. Toluene and xylene Selectivity with Polypropylene (PP). Catalytic co-pyrolysis
were the major nonoxygenated monoaromatic compounds (CCP) of lignocellulosic biomass (LB) and polypropylene
observed, irrespective of the change in the feed ratios. Olefins (PP) mixtures was carried out at three different mass ratios
were the second most abundant compounds in the TCP of LB (1:1, 1:2, and 2:1). The co-pyrolysis experiments were
and PP mixtures. The olefins in the pyrolysis liquid are repeated three times to check the consistency of the results.
branched and are mainly produced from the decomposition of Figure 6 compares the percentage yields of liquid, char, and
PP. The small changes observed in the olefin and aromatic gases at different LB to PP ratios. The liquid yield was
463 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

polymer, provides significant hydrogen to the LB and increases


the hydrocarbon yield. Hence, the liquid yield is expected to
increase at high PP proportions as observed in Figure 6. In all
the co-pyrolysis experiments, phase separation was observed in
the final liquid; one phase was rich in hydrocarbons, and the
other was rich in oxygenates (see Table S3 in the Supporting
Information for phase distribution values). Figure 7 summa-

Figure 5. Distribution of functional groups as a function of feed ratio


for thermal pyrolysis (LB and PS).

Figure 7. Distribution of functional groups as a function of feed ratio


for CCP of LB and PP mixtures.

rizes the product distribution in the CCP of LB and PP


mixtures. Increasing the PP proportion in the feed mixture
boosted the formation of useful hydrocarbons (aromatics and
olefins). The combined yield of aromatics and olefins increased
from ∼65 to ∼89%C as the LB to PP ratio varied from 2 to
0.5. The increased PP loading resulted in a decrease in
undesired char, from ∼28 to ∼13 wt % (see Figure 6). Thus,
maintaining a hydrogen-rich environment during CCP has
contributed to the conversion of biomass into more useful
hydrocarbons and lowered the char formation.
Figure 6. Product yields as a function of feed ratio for CCP of LB and 3.4.1. Effect of Feed Ratio on Aromatic and Olefin Yields.
PP mixtures. Both the aromatic and olefin yields increased at high PP
proportion in the feed. The maximum aromatic yield (∼49 %
maximum (∼43 wt %) at an LB to PP ratio of 0.5 and C) was obtained for an LB to PP ratio of 0.5, whereas the
minimum (∼35%) at a ratio of 2. This observation is in line minimum (∼32%C) was obtained for an LB to PP ratio of 1.
with the expectation since the pyrolysis of PP gives a Though the aromatic yield at an LB to PP ratio of 2 was higher
significantly higher liquid yield (∼60 wt %; see Figure 2), than the intermediate ratio, the majority of the aromatics at
and hence, the mixture with a higher proportion of PP exhibits high biomass fractions were phenolic compounds that
the maximum liquid yield. Similarly, it can be argued that originated from biomass. Though the aromatic formation can
biomass pyrolysis results in a higher char yield compared to happen in many ways, it is attributed primarily to the
PP, and therefore, the char yields in Figure 6 increase with interaction of furans (from cellulose) with PP-derived olefins.
biomass fraction in the feed. Cellulose is the principal constituent of LB and a polymer of β-
The influence of LB on the PP decomposition is explained (1→4)-D-glucopyranose units. During pyrolysis, cellulose
as follows. LB begins to degrade at a much lower temperature depolymerizes into anhydrosugars (mostly levoglucosan) and
than plastic. The thermal breakdown of plastic proceeds as a other oxygenated compounds.6,40 In the presence of an acid
radical chain process that includes steps of radical initiation, catalyst, the anhydrosugars undergo dehydration and rear-
chain propagation, and radical termination. During the initial rangement reactions to form smaller oxygenates (e.g., furans,
stage of co-pyrolysis, radicals from the LB components initiate smaller aldehydes, and water) that penetrate the catalyst pores
the chain scission of the plastic polymer chain. Consequently, and undergo a series of reactions (dehydration, dehydrogen-
the hydrogen transfer takes place from the PP chain to the LB- ation, decarboxylation, decarbonylation, oligomerization, and
derived radicals, which stabilize LB degradation products. This isomerization) to form monocyclic aromatics and olefins.2
mechanism, in turn, helps improve the hydrocarbon yield and Moreover, Diels−Alder reaction between furans and olefins
restricts char formation.38,39 PP, being a hydrogen-rich results in the formation of a Diels−Alder adduct, which, upon
464 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

dehydration, produces aromatics. The Diels−Alder reaction is predominantly produces branched olefins. It is reported that
the most desired in biomass co-pyrolysis since it produces C6 the branched olefins are less reactive than linear α-olefins due
ring compounds in a single step (see Figure 8). It is reported to steric hindrance.16 Branched alkanes were observed with a
<7%C yield in all the co-pyrolysis experiments; these are
envisaged to be formed by hydrogenation of the branched
olefins through H2 released during the aromatization process.
Commonly observed alkanes were 1,3,5-trimethylcyclohexane
and 2,6-dimethylnonane.
3.4.2. Effect of Feed Ratio on Aromatic Hydrocarbon
Selectivity. Figure 9 shows the selectivity of the aromatic

Figure 8. Diels−Alder reaction between furans and olefins in the


presence of spent FCC catalyst.

that the furan acts as a conjugated diene and the olefin as a


dienophile.3,22,41 Furans and olefins are the key reactants for
the aromatic formation in co-pyrolysis, and a low abundance of
either one of them would significantly vary the aromatic yield.
Co-feeding PP, therefore, ensures the presence of olefins in the
pyrolysis reactions, where it aids both in the aromatic
formation and in the stabilization of LB-derived oxygenates.
In addition to their contribution to aromatic formation, the
PP-derived olefins also undergo a series of reactions
(oligomerization, cyclization, and aromatization) within the
catalyst pores.16 However, the rate of olefins to aromatic
formation through PP depolymerization is slow compared to
the LB pyrolysis reactions. Figure 9. Aromatic selectivity as a function of feed ratio for CCP of
Char formation is a significant competing reaction that LB and PP mixtures.
hinders the aromatic formation. It can happen in many ways
via polymerization of degraded cellulosic fragments or compounds, which were the most desired in this study.
polymerization of furans and aromatics.21,42 LB is rich in Toluene, xylene, alkylbenzenes, and phenols (cresols, benze-
carbon relative to hydrogen, and most of the hydrogen in LB is nediols, and xylenols) were the main aromatic compounds
bound to oxygen in the form of a hydroxyl group rather than detected. The selectivity for toluene peaked (∼24%C) at an LB
carbon. During LB pyrolysis, most of this hydrogen is removed to PP ratio of 1. Selectivity toward xylene increased from ∼21
as water, making it more hydrogen-deficient. As a result, the to ∼35%C with an increase in the PP loading in the feed. A
reaction mechanism shifts toward the char formation due to similar behavior was observed by Cheng et al.,46 who reported
the lack of hydrogen availability for hydrocarbon formation.5,43 that furan and propylene produced a significant amount of
The decrease in aromatic yields at high biomass fractions is xylenes in the presence of the HZSM-5 catalyst. Although the
due to the coverage of the active surface area of the catalyst by overall aromatic selectivity increased with a higher PP content
char formation. Co-pyrolysis at high and intermediate biomass in the feed, the combined yield of toluene and xylene was
fractions produce more sugars (especially levoglucosan). The maximum (∼54%C) at an LB to PP ratio of 1. This value
sugar molecules cannot diffuse readily into the catalyst pores suggests that there exists an optimal LB to PP ratio at which
owing to their large molecular size.44,45 They dehydrate on the maximum valuable aromatics are formed. Lignin contains
external active surface of the catalyst to form furans and other several methoxy groups in its structure, resulting in the
smaller oxygenates. The furans have to diffuse through the generation of methyl groups at high temperatures.47 The
catalyst pores to further get transformed to aromatics. absence of benzene and high selectivity for xylenes at high PP
However, as time proceeds, char formation happens through fractions signify that the methyl groups from lignin react with
the dehydration of accumulated sugar molecules. The char benzene to form xylene. In addition to xylene and toluene,
buildup on the external surface of the catalyst restricts the alkylbenzenes such as 3-ethyltoluene, trimethylbenzenes, and
passage of smaller molecules like furans into the catalyst pores their isomers were found substantially at all feed ratios.
hampering both the aromatic formation and deoxygenation However, their trend was not monotonous with the changes in
reactions. Therefore, high amounts of furans (oxygenates) at a the feed ratio.
high biomass loading (LB to PP ratio of 2) are observed, as Other monoaromatic compounds like phenols (including
seen in Figure 7. phenols, cresols, and their isomers, benzenediols and their
Similar to aromatics, the olefin content increased with isomers, and xylenols) are found abundantly in the bio-oil.
increasing PP fraction in the feed. The highest yield of ∼40%C Their selectivity was ∼43 %C at high biomass ratios and
and the lowest yield of ∼19%C were observed at LB to PP decreased to ∼27 %C at a low biomass fraction. Previous
ratios of 0.5 and 2, respectively. The olefins were mainly in the studies have reported that the phenols are formed either by the
range of C7−C12, and C9 fraction was the most dominant. breakdown of ether linkages in the main lignin chain or
Most of the unreacted olefins were branched olefins. This is through the oxidation of aromatic intermediates.22,25 Lignin is
because PP is a branched polymer, and its thermal breakdown one of the major components of lignocellulosic biomass and is
465 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

itself a cross-linked phenolic polymer that contains alkoxy, biomass loading, the absence of furans indicates that the
aryloxy, and phenolic subunits. The major pyrolysis products aromatic formation was greatly enhanced through the Diels−
of lignin are phenols, guaiacols, and syringols. It has been Alder mechanism, as discussed in the previous subsection.
recently reported that guaiacol acts as a methylating agent and 3.4.4. Effect of Feed Ratio on Olefin Distribution. Figure
forms benzenediols, hydroxybenzene, and cresols as final 10 depicts the olefin distribution during CCP of LB and PP
products in the presence of the H-USY catalyst.48 Our
observations are consistent with this reported trend as
substantial quantities of phenols were formed at a high
biomass fraction in the feed.
Lignin-derived phenols may also be the precursors for the
formation of PAHs (polycyclic aromatic hydrocarbons) and
char. Phenols are relatively thermally stable and are less
reactive compared to other molecules. However, they tend to
dehydrate and polymerize in the presence of an acid catalyst.
Recent studies report that PAH formation is mainly due to the
combination of biomass-derived oxygenates with monoaro-
matics, viz., secondary reactions of pyrolysis products.9,49 In
our study, the PAH selectivity was ∼5% C at the lowest LB to
PP ratio. However, it was negligible for the other two ratios
due to coverage of the active catalyst surface area by char, as
mentioned in section 3.4, at high biomass loading, which
inhibits PAH formation owing to the inaccessibility of the
pores for the dehydration and polymerization reactions.
3.4.3. Effect of Feed Ratio on Oxygenate Distribution. Figure 10. Olefin distribution as a function of feed ratio for CCP of
CCP of LB and PP mixtures resulted in oxygenated LB and PP.
compounds like anhydrosugars, carboxylic acids, aldehydes,
ketones, alcohols, and esters. The cellulose and hemicellulose mixtures at different mass fractions. The olefins were widely
are the primary sources for their formation. Table 2 shows the distributed ranging from C7 to C12 in all the experiments.
Majority of the unreacted olefins belong to C9, which have
Table 2. %C Yields of Oxygenated Compounds in Catalytic branching in their structures, for example, (Z)-2,4-dimethyl-1-
Co-pyrolysis of LB and PP heptene, (Z)-2,3-dimethyl-3-heptene, and (Z)-2,6-dimethyl-3-
LB/PP ratio heptene; they constituted ∼50% of the olefin composition at
all the feed ratios. Li et al.16 reported that branched olefins do
compounds 0.5 1 2
not interact appreciably in the Diels−Alder reaction due to
anhydrosugars 0.76 0 0.95 their restricted mobility into the catalyst pores. Apart from
furans 0 9.87 8.37 branched olefins, C7 cyclic olefins like 1,3-cycloheptadiene
carboxylic acids 1.07 4.65 0.66 were also found in abundance at higher and intermediate PP
esters 1.92 1.68 4.54 loading. They represent ∼36% of the olefins composition at an
ethers 0 0 4.77 LB to PP ratio of 0.5. This suggests that cyclic olefins are solely
ketones 0 5.97 4.21 formed from the PP decomposition and not from the LB
alcohols 0.48 4.38 5.79 components. The presence of furans in the liquid suggests that
total 4.23 26.54 29.29 furans interact with the smaller olefins only which are in cis-
conformation.16 The possible reaction mechanisms during
%C selectivity of oxygenated compounds; the major CCP of LB and PP are depicted in Figure 11, which shows the
compounds are furans, hydroxyl, carboxyl, and carbonyl different routes for formation of furans (e.g., directly from
group compounds. As expected, the amount of oxygenated hemicellulose pyrolysis or by dehydration of cellulose pyrolysis
compounds diminished as the PP portion in the feed mixture products), aromatic formation from olefins derived through
increased since oxygenates are mainly formed from biomass. In cracking of PP or pyrolysis of lignin, etc.; it may be noted that
addition to dehydration, oxygen is also removed by decarbon- these reaction pathways are already described previously as
ylation and decarboxylation reactions in the form of CO and needed in this section.
CO2. The significant decrease in carboxyl and carbonyl groups 3.4.5. Hydrogen Utilization. To evaluate the effective
at high PP loading may be mainly attributed to the better hydrogen utilization at various feed ratios, the (H/C)eff ratios
catalytic activity to disrupt the −COOH and CO bonds to were calculated and compared with the bio-oil obtained from
produce CO2, CO, and hydrocarbons as final products.45 the thermal and catalytic pyrolysis of biomass. The (H/C)eff
The presence of significant amounts of unconverted furans ratios in Figure 12 were calculated using the results of the
and olefins at high LB loading reveals that char buildup on the CHNS analysis of the liquid products and eq 7. As expected,
catalyst surface had a significant impact on the interaction the (H/C)eff ratio in the product from catalytic pyrolysis is
between the two classes of compounds. The char formation, in higher than thermal pyrolysis; therefore, it can be inferred that
turn, influenced the aromatic yield as there was less provision addition of the spent FCC catalyst has led to an improvement
for the molecules to access the acidic sites of the catalyst. in the bio-oil quality. It can be observed from Figure 12 that
Among the oxygen-containing aromatics, phenols contributed the CCP product values at LB to PP ratios of 0.5 and 1 are
to around ∼43% of their composition at high biomass loading; significantly higher than the products of thermal (TP) or
these are primarily the degradation products of lignin. At low catalytic (CP) pyrolysis. The trend indicates that hydrogen
466 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

Figure 11. Possible reaction mechanisms during CCP of LB and PP mixtures with spent FCC.

For co-pyrolysis experiments, CH4 was the most dominating


compound during TCP, whereas carbon oxides were the most
abundant gases during CCP. For TCP, the highest CH4 yield
(∼61%) was observed at an LB/PP ratio of 0.5, while the
corresponding value was ∼13% at an LB/PP ratio of 2 for
CCP. The H2 yield was almost the same for TCP and CCP
runs, with the yield values ranging from ∼21 to 28%. The peak
H2 value (∼26 to 27%) was observed at an LB/PP ratio of 0.5
in the case of the co-pyrolysis experiments. With the addition
of the catalyst, the yield of carbon oxides enhanced greatly with
a simultaneous decrease in CH4 yields. For CCP, the carbon
oxides yields were in the range of ∼63−75% with peak value
(∼74%) at an LB/PP ratio of 1. In the case of TCP, the yield
range of the carbon oxides was only 13 to 37% with peak value
(∼37%) at an LB/PP ratio of 1. These results suggest that the
catalyst addition upgraded the bio-oil by disrupting the
carbonyl (CO) and carboxyl (COOH) functional groups
Figure 12. (H/C)eff ratio as a function of feed composition for
pyrolysis of LB and PP mixtures from biomass and removing the oxygen in the form of CO and
CO2.
3.6. Effect of Feed Composition on Yields and
from plastic was effectively utilized at LB to PP ratios of 0.5 Selectivity with Polystyrene (PS). Figure 14 compares the
and 1, which implies that CCP of LB and PP improved the H/ product yields for CCP of lignocellulosic biomass (LB) and
C ratio of the bio-oil and the H2 from plastic helped in polystyrene (PS) at mixing ratios of 1:1, 1:2, and 2:1. As
lowering the oxygen content of the bio-oil. The low EHI value expected, the addition of PS to biomass improved the liquid
at the highest biomass loading (feed ratio of 2) is close to that yield since the stand-alone pyrolysis of PS gives a significant
of the product from catalytic pyrolysis of biomass; this is a amount of liquid (>80 wt %; see Figure 4) compared to
result of the higher fraction of aromatics in the bio-oil at an LB biomass. The bio-oil yield was minimum (∼46%) at a ratio of
to PP ratio of 2, which inherently have a lower H/C ratio (e.g., 2 and increased to a maximum (∼65%) with the increase in
1 for benzene) compared to olefins and alkanes. the PS content in the feed at an LB to PS ratio of 0.5. Also, the
3.5. Effect of Feed Ratio on Pyrolysis Gas Distribution pyrolysis of PS produces significantly lower residue amounts
in CCP of LB-PP. Figure 13 represents the pyrolysis gas than biomass. Hence, the char yields in CCP of LB and PS can
distribution as a function of feed ratio. H2, CH4, and carbon be mainly attributed to LB, and it varies between ∼17 to ∼26%
oxides (CO and CO2) were the main gases observed during as the LB fraction in the feed increases.
TP, TCP, and CCP experiments. In LB pyrolysis, carbon 3.6.1. Comparison of Bio-Oil Component Distribution in
oxides were the dominant gases comprising ∼56% of the total CCP of LB and PS Mixtures. Figure 15 summarizes the product
gas yield, followed by CH4 and H2 with yield values of ∼24 and carbon yields from catalytic co-pyrolysis of LB and PS
∼20%, respectively. In thermal pyrolysis of PP, CH4 and H2 mixtures. The aromatic selectivity was almost the same at all
were the only observed gases with yield values of ∼62 and the mixing ratios, and the values are in the range of ∼79−82%
∼38%, respectively. No carbon oxides were detected as PP has C. The high aromatic content in the liquid is primarily due to
only carbon and hydrogen in its structure. PS, which is an aromatic polymer and its pyrolysis produces
467 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

Figure 13. Gas distribution as a function of feed composition for thermal, thermal co-pyrolysis (TCP), and catalytic co-pyrolysis (CCP) of LB and
PP mixtures.

Figure 15. Distribution of functional groups as a function of feed ratio


Figure 14. Product yields as a function of feed ratio for CCP of LB for CCP of LB and PS mixtures.
and PS mixtures.
indene derivatives were the common products during LB-PS
aromatic compounds. The degradation products of lignin also co-pyrolysis. However, their selectivity varied significantly
contribute to aromatics in the pyrolysis liquid. The aromatic among the three ratios. The % carbon yields of the aromatic
formation through the Diels−Alder mechanism is not possible hydrocarbons during CCP of LB-PS mixtures for the three
as PS is not capable of producing smaller olefins.33 Similarly, ratios are presented in Table 3.
there was no significant change in the oxygenate content in the Styrene was the most abundant aromatic product from CCP
liquid with the changing feed ratio. The oxygenate values were of LB-PS mixtures. Its composition was maximum (∼27%C) at
found to be in the range of ∼18−20%C. It can be inferred that an LB to PS ratio of 1 and minimum (∼16%C) at a ratio of 2.
there were no significant interactions between the PS and LB Polystyrene is a polymer of styrene, and the β-scission of the
components as the aromatic rings are more stable compounds C−C bond in its main chain produces styrene. Also, the
compared to olefins. presence of an acidic catalyst accelerates the disruption of the
3.6.2. Effect of Feed Ratio on the Aromatic Hydrocarbon C−C bonds in the PS polymer chain.50,51 The lower selectivity
Distribution. The aromatics were the desired compounds of of styrene monomer at higher biomass loading is attributed to
interest in this study. Styrene, ethylbenzene, α-methyl styrene, either limited catalytic activity or further styrene monomer
styrene dimers, phenols, alkylbenzenes, naphthalene, and conversion to compounds such as ethylbenzene, styrene
468 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

Table 3. Aromatic Selectivity as a Function of Feed Ratio ethylbenzene, which is reported to occur in the presence of an
for CCP of LB and PS Mixtures acid catalyst.50 The ethylbenzene formation was in the range of
∼14 to ∼17%C as the feed ratio changed.
LB/PS ratio
Alkylaromatics such as α-methylstyrene, α-ethylstyrene, and
aromatic hydrocarbons 0.5 1 2 isopropylbenzene were also observed at all the feed ratios.
styrene 24.02 26.97 15.79 Their formation can be mainly attributed to β-scission and
ethylbenzene 13.92 16.93 15.28 hydrogen transfer to PS.44
styrene dimers 13.63 15.03 13.86 Styrene dimers were also observed in the liquid product at
benzene derivatives 7.67 4.80 14.30 all feed ratios; they are styrene oligomers typically formed by
alkyl styrene 7.75 13.15 1.80
the back-biting reaction of styrene monomers through the free
phenols 6.73 8.12 5.73
radical mechanism followed by β-scission during depolymeri-
PAHs (naphthalene and its derivatives, indene 26.29 15.0 32.88
and its derivatives, and multiring aromatics zation of the styrene polymer.52 The percentage selectivity of
having more than two rings) styrene dimers was ∼14% C on average. The observed styrene
dimers include 2-phenyltetralin, diphenylmethane, 1,3-diphe-
derivatives, and styrene dimers. The formation of ethylbenzene nylpropane, 1,2-diphenylpropane, 1,3-diphenyl-1-butene, 1,2-
suggests that hydrogen transfer happened during co-pyrolysis. diphenylethylene, and 1,5-diphenyl-1-pentene. Among all the
Figure 16 illustrates the reaction of hydrogenation of styrene to styrene dimers, 1,3-diphenylpropane and 1,3-diphenyl-1-
butene were predominant in all the experimental runs. The
maximum yield for 1,3-diphenylpropane (∼5%) was obtained
at low PS loading, and the maximum yield (∼5%) for 1,3-
diphenyl-1-butene was obtained at high PS loading. 1,3-
Diphenylpropane is formed by the reaction between the
styrene radical and benzyl radical, while 1,3-diphenyl-1-butene
is formed by the combination of two styrene free radicals
followed by dehydrogenation.33 The possible reaction path-
Figure 16. Hydrogenation of styrene to ethylbenzene. ways for the formation of various styrene dimers are depicted
in Figure 17.

Figure 17. Reaction pathways for styrene dimer formation during CCP of LB and PS mixtures.33

469 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

Figure 18. Mechanism of PAH formation during lignin depolymerization.58

Other compounds like benzene derivatives were also found The high PAH selectivity and low phenolics yield at high
in the bio-oil, especially in a higher percentage (∼14%) at a biomass loading indicates a greater possibility of phenols acting
feed ratio of 2. These benzene derivatives include iso- as potential precursors for PAH formation. The major phenolic
propylbenzene, propylbenzene, 2-propenylbenzene, 1-prope- compounds identified include phenols, cresols (and their
nylbenzene, cyclopropylbenzene, biphenyl, terphenyl, etc. isomers), and benzenediols (and their isomers), which are of
Apart from the desired monoaromatics, PAHs, which are lignin origin formed via breakage of weak α-ether and β-ether
extremely undesirable in the bio-oil, are also found. They are bonds.59 In the presence of a catalyst, they actively undergo a
highly susceptible to coke formation in the presence of an acid series of dehydration and polymerization reactions to form
catalyst. In this study, PAHs are categorized as naphthalene PAH and char.56 However, the actual mechanism of PAH
and its derivatives, indene and its derivatives, and multiring formation is quite elusive and yet to be understood completely.
aromatic components having more than two rings. The total 3.6.3. Effect of Feed Ratio on Oxygenate Distribution.
PAH selectivity was maximum (∼33%) at high LB loading and Table 4 shows the %C selectivity of oxygenated compounds in
minimum (∼15%) at an equal feed ratio. Among the PAH
observed, two-ring PAH, naphthalenes and indenes, dominated Table 4. %C Yields of Oxygenated Compounds in Catalytic
in all the liquid samples. Phenanthrene and fluorene were the Co-pyrolysis of LB and PS
second most detected three-ring PAHs. Among the two-ring LB/PS ratio
PAHs, the maximum selectivity for naphthalene derivatives
compounds 0.5 1 2
(∼14%) was observed at high and low PS loadings, while the
maximum for indene derivatives (∼12%) was observed at high anhydrosugars 3.46 9.48 1.30
biomass loading. Both the naphthalene and indene derivatives furans 1.35 7.97 3.30
exhibited a lower selectivity at intermediate mixing ratios (∼7 carboxylic acids/esters 2.31 1.54
and 5%, respectively). aldehydes/ketones 9.65 2.57 13.32
Since both PS and lignin are aromatic in nature, there is a alcohols 1.64 0.85 0.40
possibility of PAH formation from their degradation products. total 18.40 20.87 19.86
Monocyclic aromatics combine with oxygenated fragments
from biomass to form naphthalenes and other higher PAHs.53 the bio-oil. Anhydrosugars, furans, aldehydes, and ketones
It is reported that styrene and benzene also contribute to two- were the majority of oxygenates observed in all the feed ratios;
ring PAHs, and the presence of a catalyst can enhance their these compounds are mainly the defragmented products of
formation.54,55 Styrene monomers via intramolecular hydrogen cellulose and hemicellulose. The oxygenate content almost
transfer and β-scission reactions also lead to PAH formation.56 remained the same (∼18−21%) at all the feed ratios,
Among naphthalene derivatives, 2-phenylnaphthalene was indicating that the PS and LB components did not have
observed at all the LB to PS ratios (∼8%). Though benzene significant interactions. The low amounts of sugars and furans
was not observed in the final liquid product during co-pyrolysis and high PAH content (see Table 4) at high and low biomass
of biomass and PS, the presence of indene derivatives indicates loading reveal that these oxygenates contributed to PAH
the formation of benzene and its cyclization in the presence of formation.53
the catalyst to produce indene and its derivatives.57 The 3.6.4. Hydrogen Utilization. To evaluate the effective
presence of benzene derivatives also supports the formation of hydrogen utilization between LB and PS, a (H/C)eff analysis
benzene during co-pyrolysis. Further, both the naphthalene was performed similar to that of PP in section 3.4.5. Most of
and indene derivatives had methyl groups attached to the ring the studies in the literature indicated that PS and biomass did
and can be attributed to the generation of methyl groups not exhibit significant interactions during catalytic pyroly-
during lignin depolymerization. The presence of phenols, sis.32,33,60 In our study, the (H/C)eff values of the co-pyrolysis
cresols, and catechols in the bio-oil indicate that the secondary liquid almost remained the same at all feed ratios (see Figure
pyrolysis reactions of lignin (demethoxylation, dehydroxyla- 19), which signify that PS and LB had minimum interactions
tion, and demethylation) occurred during co-pyrolysis.58 The during co-pyrolysis.
mechanism of PAH formation during lignin depolymerization Figure 19 also shows that the CCP product (H/C)eff values
is shown in Figure 18. are higher than the products of thermal (TP) or catalytic
470 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

pyrolysis of PS. The carbon oxides are mainly the products of


LB decomposition as PS is an aromatic polymer consisting of
carbon and hydrogen only in its structure. The percentage of
carbon oxides varied in the range of ∼56−73% across all the
experiments in Figure 20. These results (particularly carbon
oxide yields) further justify that LB and PS did not have
significant interactions either in TCP or CCP.

4. CONCLUSIONS
In this study, experiments on catalytic co-pyrolysis of
lignocellulosic biomass and plastics (PP and PS) with spent
FCC catalyst were performed at three different mixing ratios to
investigate the effect of plastic composition on the product
yields and selectivity. The increase in plastic proportion in the
feed enhanced the liquid yield. The oxygenate content also
Figure 19. (H/C)eff ratio as a function of feed composition for reduced to a significant extent from 30 to 5%C when the
pyrolysis of LB and PS mixtures. plastic proportion was increased from 33 to 67 wt %, especially
with PP (Figure 7). The results suggest that decarboxylation
pyrolysis (CP) at mixing ratios where the plastic proportion is and decarbonylation reactions were the primary deoxygenation
low. At the highest PS loading, the (H/C)eff value is close to pathways during CCP using the spent FCC catalyst. The
that of catalytic pyrolysis, possibly due to the formation of
maximum yield of monoaromatics and olefins was ∼73% when
monoaromatics. At other PS loading ratios, the proportion of
PP was used as a co-reactant at an LB to PP ratio of 0.5. Using
alkyl styrenes is higher, which results in (H/C)eff values greater
than one. PS as a co-feed produced the maximum aromatic yield of
3.7. Effect of Feed Ratio on Pyrolysis Gas Distribution ∼82% at the same feed ratio. However, the more valuable
in CCP of LB-PS. Figure 20 shows the pyrolysis gas chemicals like toluene, xylene, etc., were found abundant for
PP, and useful aromatics like styrene, ethylbenzene, etc., were
abundantly formed for PS at equal mixing ratios. This suggests
that the optimum feed conditions would tailor the reaction
mechanism toward the formation of more valuable chemicals
due to the presence of active acidic sites of the spent FCC
catalyst. Further, since the best performance in terms of
aromatic yields was obtained at a feed ratio of 0.5, it can be
concluded that having a large excess of plastic in the co-
pyrolysis feed mixture is not beneficial; this is important as
obtaining greater amounts of waste plastic compared to
biomass is challenging. The waste agricultural residue or
biomass generation is much higher than waste plastic
generation. The (H/C)eff ratio of the final liquid products in
the case of PP was comparatively higher than the products of
catalytic pyrolysis of biomass, which suggests oxygen reduction
during CCP process; this is also expected to increase the
calorific value of the oil formed. However, in the case of PS,
the (H/C)eff results show that PS did not have significant
interactions with LB. These results suggest that the product
distribution in CCP of lignocellulosic biomass and plastic
Figure 20. Gas distribution as a function of feed composition for mixtures strongly depends on the feed composition. Also, co-
thermal, thermal co-pyrolysis (TCP), and catalytic co-pyrolysis pyrolysis of a H2-rich feedstock (plastic) and a H2-deficient
(CCP) of LB and PS mixtures.
feed (biomass) offers synergy in improving the (H/C)eff of the
bio-oil formed in the presence of spent FCC catalyst. Future
compositions as a function of feed ratio for TP of LB and
studies could consider optimization of the reaction conditions
PS and co-pyrolysis (thermal and catalytic) of LB and PS
mixtures. H2, CH4, and carbon oxides (CO and CO2) were the by investigating the effects of temperature, heating rate, catalyst
most abundant gases in all the experimental runs. The loading and residence time together. A continuous process
percentage of H2 in LB pyrolysis was ∼20%, whereas this rather than a batch process might be more amenable and could
value ranged from ∼11 to 14% and ∼8 to 18% for TCP and be taken up as another future work using auger reactors. This
CCP, respectively. process helps in valorization of waste plastics and agricultural
The percentage CH4 yield was maximum (∼24%) for LB residues to make chemicals and in waste management.
pyrolysis and TCP (∼23%) at a feed ratio of 0.5. Carbon However, owing to the logistics involved in the collection,
oxides were the most dominant compounds during TP of LB, transport, and storage of these waste resources, it is advisable
TCP, and CCP of LB-PS mixtures, irrespective of the feed to have small-scale and distributed CCP units, which can be
ratios. There was no significant gas yield in the case of thermal mobile and modular for easy scale-up of the process.
471 DOI: 10.1021/acs.energyfuels.9b03135
Energy Fuels 2020, 34, 460−473
Energy & Fuels Article


*
ASSOCIATED CONTENT
S Supporting Information
(12) Alonso, D. M.; Bond, J. Q.; Dumesic, J. A. Catalytic conversion
of biomass to biofuels. Green Chem. 2010, 12, 1493−1513.
(13) Vispute, T. P.; Zhang, H.; Sanna, A.; Xiao, R.; Huber, G. W.
The Supporting Information is available free of charge at Renewable Chemical Commodity feedstocks from Integrated
https://pubs.acs.org/doi/10.1021/acs.energyfuels.9b03135. Catalytic Processing of Pyrolysis Oils. Science 2010, 330, 1222−1227.
Table summarizing the use of catalyst in co-pyrolysis of (14) Zhang, Q.; Chang, J.; Wang, T.; Xu, Y. Review of biomass
plastics and biomass; proximate analysis of the feed pyrolysis oil properties and upgrading research. Energy Convers.
components; tables showing the % weight distribution of Manage. 2007, 48, 87−92.
organic and aqueous phases in the bio-oil and % yields of (15) Zhang, X.; Lei, H.; Zhu, L.; Zhu, X.; Qian, M.; Yadavalli, G.;
products (liquids, gases, and char) in the catalytic Wu, J.; Chen, S. Thermal behavior and kinetic study for catalytic co-
pyrolysis of LB, PP, and PS; and figure showing the pyrolysis of biomass with plastics. Bioresour. Technol. 2016, 220, 233−
DTG curves for LB, PP, and PS (PDF) 238.


(16) Li, X.; Li, J.; Zhou, G.; Feng, Y.; Wang, Y.; Yu, G.; Deng, S.;
Huang, J.; Wang, B. Enhancing the production of renewable
AUTHOR INFORMATION petrochemicals by co-feeding of biomass with plastics in catalytic
Corresponding Author fast pyrolysis with ZSM-5 zeolites. Appl. Catal., A 2014, 481, 173−
*E-mail: s.srinivas@iitb.ac.in. 182.
(17) Zhang, H.; Nie, J.; Xiao, R.; Jin, B.; Dong, C.; Xiao, G. Catalytic
ORCID Co-pyrolysis of Biomass and Different Plastics (Polyethylene,
Seethamraju Srinivas: 0000-0003-0505-1029 Polypropylene, and Polystyrene) To Improve Hydrocarbon Yield in
Notes a Fluidized-Bed Reactor. Energy Fuels 2014, 28, 1940−−1947.
The authors declare no competing financial interest. (18) Han, B.; Chen, Y.; Wu, Y.; Hua, D.; Chen, Z.; Feng, W.; Yang,


M.; Xie, Q. Co-pyrolysis behaviors and kinetics of plastics−biomass
ACKNOWLEDGMENTS blends through thermogravimetric analysis. J. Therm. Anal. Calorim.
2014, 115, 227−235.
The authors thank IRCC, IIT Bombay for the seed grant (19) Li, X.; Zhang, H.; Li, J.; Su, L.; Zuo, J.; Komarneni, S.; Wang, Y.
project funding provided to S.S. and MHRD, India for Improving the aromatic production in catalytic fast pyrolysis of
providing the scholarship to K.P.K. during his graduate cellulose by co-feeding low-density polyethylene. Appl. Catal., A 2013,
studentship. The authors also thank SAIF, IIT Bombay for 455, 114−121.
providing the analytical facilities used in this work. The authors (20) Geyer, R.; Jambeck, J. R.; Law, K. L. Production, use, and fate
are also thankful to the anonymous reviewers whose comments of all plastics ever made. Sci. Adv. 2017, 3, No. e1700782.
helped in improving the quality and presentation of this work. (21) Tan, S.; Zhang, Z.; Sun, J.; Wang, Q. Recent progress of

■ REFERENCES
(1) British Petroleum, BP statistical Review of World Energy, 2019,
catalytic pyrolysis of biomass by HZSM-5. Chin. J. Catal. 2013, 34,
641−650.
(22) Mullen, C. A.; Dorado, C.; Boateng, A. A. Catalytic co-pyrolysis
68th edition, page 4, https://www.bp.com/content/dam/bp/ of switchgrass and polyethylene over HZSM-5: Catalyst deactivation
business-sites/en/global/corporate/pdfs/energy-economics/ and coke formation. J. Anal. Appl. Pyrolysis 2018, 129, 195−203.
statistical-review/bp-stats-review-2019-full-report.pdf, Last accessed (23) Zhang, H.; Xiao, R.; Nie, J.; Jin, B.; Shao, S.; Xiao, G. Catalytic
on 6th Dec 2019. pyrolysis of black-liquor lignin by co-feeding with different plastics in
(2) Mohan, D.; Pittman, C. U., Jr.; Steele, P. H. Pyrolysis of Wood/ a fluidized bed reactor. Bioresour. Technol. 2015, 192, 68−74.
Biomass for Bio-oil: A Critical Review. Energy Fuels 2006, 20, 848− (24) Hong, Y.; Lee, Y.; Rezaei, P. S.; Kim, B. S.; Jeon, J.-K.; Jae, J.;
889. Jung, S.-C.; Kim, S. C.; Park, Y.-K. In-situ catalytic co-pyrolysis of
(3) Huber, G. W.; Iborra, S.; Corma, A. Synthesis of Transportation cellulose and polypropylene over desilicated ZSM-5. Catal. Today
Fuels from Biomass: Chemistry, Catalysts, and Engineering. Chem. 2017, 293-294, 151−158.
Rev. 2006, 106, 4044−4098. (25) Zhao, Y.; Wang, Y.; Duan, D.; Ruan, R.; Fan, L.; Zhou, Y.; Dai,
(4) World Bioenergy Association WBA Global Energy Statistics;, L.; Lv, J.; Liu, Y. Fast microwave-assisted ex-catalytic co-pyrolysis of
2018, https://worldbioenergy.org/uploads/WBA%20GBS%202017_ bamboo and polypropylene for bio-oil production. Bioresour. Technol.
hq.pdf, Last accessed 3rd August 2019. 2018, 249, 69−75.
(5) Bridgwater, A. V. Review of fast pyrolysis of biomass and product (26) Qi, P.; Chang, G.; Wang, H.; Zhang, X.; Guo, Q. Production of
upgrading. Biomass Bioenergy 2012, 38, 68−94. aromatic hydrocarbons by catalytic co-pyrolysis of microalgae and
(6) French, R.; Czernik, S. Catalytic pyrolysis of biomass for biofuels polypropylene using HZSM-5. J. Anal. Appl. Pyrolysis 2018, 136, 178−
production. Fuel Process. Technol. 2010, 91, 25−32. 185.
(7) Huber, G. W.; Corma, A. Synergies between Bio- and Oil (27) Suriapparao, D. V.; Boruah, B.; Raja, D.; Vinu, R. Microwave
Refineries for the Production of Fuels from Biomass. Angew. Chem.
assisted co-pyrolysis of biomasses with polypropylene and polystyrene
Int. Ed. 2007, 46, 7184−7201.
for high quality bio-oil production. Fuel Process. Technol. 2018, 175,
(8) Czernike, S.; Bridgwater, A. V. Overview of Applications of
Biomass Fast Pyrolysis Oil. Energy Fuels 2004, 18, 590−598. 64−75.
(9) Carlson, T. R.; Cheng, Y.-T.; Jae, J.; Huber, G. W. Production of (28) Chi, Y.; Xue, J.; Zhuo, J.; Zhang, D.; Liu, M.; Yao, Q. Catalytic
green aromatics and olefins by catalytic fast pyrolysis of wood co-pyrolysis of cellulose and polypropylene over all-silica mesoporous
sawdust. Energy Environ. Sci. 2011, 4, 145−161. catalyst MCM-41 and Al-MCM-41. Sci. Total Environ. 2018, 633,
(10) Kabir, G.; Hameed, B. H. Recent progress on catalytic pyrolysis 1105−1113.
of lignocellulosic biomass to high-grade bio-oil and bio-chemicals. (29) Sanahuja-Parejo, O.; Veses, A.; Navarro, M. V.; López, J. M.;
Renewable Sustainable Energy Rev. 2017, 70, 945−967. Murillo, R.; Callén, M. S.; García, T. Drop-in biofuels from the co-
(11) Gunawan, R.; Li, X.; Lievens, C.; Gholizadeh, M.; Chaiwat, W.; pyrolysis of grape seeds and polystyrene. Chem. Eng. J. 2019, 120246.
Hu, X.; Mourant, D.; Bromly, J.; Li, C.-Z. Upgrading of bio-oil into (30) Van Nguyen, Q.; Choi, Y. S.; Choi, S. K.; Jeong, Y. W.; Kwon,
advanced biofuels and chemicals. Part I. Transformation of GC- Y. S. Improvement of bio-crude oil properties via co-pyrolysis of pine
detectable light species during the hydrotreatment of bio-oil using sawdust and waste polystyrene foam. J. Environ. Manage. 2019, 237,
Pd/C catalyst. Fuel 2013, 111, 709−717. 24−29.

472 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473
Energy & Fuels Article

(31) Shadangi, K. P.; Mohanty, K. Co-pyrolysis of Karanja and Niger (51) Ukei, H.; Hirose, T.; Horikawa, S.; Takai, Y.; Taka, M.; Azuma,
seeds with waste polystyrene to produce liquid fuel. Fuel 2015, 153, N.; Ueno, A. Catalytic degradation of polystyrene into styrene and a
492−498. design of recyclable polystyrene with dispersed catalysts. Catal. Today
(32) Rutkowski, P.; Kubacki, A. Influence of polystyrene addition to 2000, 62, 67−75.
cellulose on chemical structure and properties of bio-oil obtained (52) Yang, M.; Shibasaki, Y. Mechanisms of Thermal Degradation of
during pyrolysis. Energy Convers. Manage. 2006, 47, 716−731. Polystyrene, Polymethacrylonitrile, and Their Copolymers on Flash
(33) Hassan, E. B.; Elsayed, I.; Eseyin, A. Production high yields of Pyrolysis. J. Polym. Sci., Part A: Polym. Chem. 1998, 36, 2315−2330.
aromatic hydrocarbons through catalytic fast pyrolysis of torrefied (53) Liu, W.-J.; Li, W.-W.; Jiang, H.; Yu, H.-Q. Fates of Chemical
wood and polystyrene. Fuel 2016, 174, 317−324. Elements in Biomass during Its Pyrolysis. Chem. Rev. 2017, 117,
(34) Chiranjeevi, T.; Pragya, R.; Gupta, S.; Gokak, D. T.; Bhargava, 6367−6398.
S. Minimization of Waste Spent Catalyst in Refineries. Procedia (54) Zhao, L.; Zhao, Y.; Nan, H.; Yang, F.; Qiu, H.; Xu, X.; Cao, X.
Environ. Sci. 2016, 35, 610−617. Suppressed formation of polycyclic aromatic hydrocarbons (PAHs)
(35) World Agricultural Production, Circular series, WAP 11−19 during pyrolytic production of Fe-enriched composite biochar. J.
November 2019, United States Department of Agriculture, https:// Hazard. Mater. 2019, 121033.
apps.fas.usda.gov/psdonline/circulars/production.pdf, Last accessed (55) Yu, Y.; Li, X.; Su, L.; Zhang, Y.; Wang, Y.; Zhang, H. The role
on 30th Nov 2019. of shape selectivity in catalytic fast pyrolysis of lignin with zeolite
(36) Agricultural and Processed Food Products Export Development catalysts. Appl. Catal., A 2012, 447-448, 115−123.
Authority (APEDA), India https://apeda.gov.in/apedawebsite/ (56) Kelkar, S.; Saffron, C. M.; Andreassi, K.; Li, Z.; Murkute, A.;
HACCP/2018_Groundnut_Survey_Report.pdf, Last accessed on Miller, D. J.; Pinnavaia, T. J.; Kriegel, R. M. A survey of catalysts for
aromatics from fast pyrolysis of biomass. Appl. Catal., B 2015, 174-
30th Nov 2019.
175, 85−95.
(37) Chen, N. Y.; Walsh, D. E.; Koenig, L. R., Fluidized-Bed
(57) Ojha, D. K.; Vinu, R. Resource recovery via catalytic fast
Upgrading of Wood Pyrolysis Liquids and Related Compounds, Pyrolysis
pyrolysis of polystyrene using zeolites. J. Anal. Appl. Pyrolysis 2015,
Oils from Biomass, ACS Symposium Series; American Chemical Society:
113, 349−359.
Washington, DC, 1988, pp. 277−289. (58) Zhou, H.; Wu, C.; Onwudili, J. A.; Meng, A.; Zhang, Y.;
(38) Sharypov, V. I.; Beregovtsova, N. G.; Kuznetsov, B. N.; Williams, P. T. Polycyclic Aromatic Hydrocarbon Formation from the
Membrado, L.; Cebolla, V. L.; Marin, N.; Weber, J. V. Co-pyrolysis of Pyrolysis/ Gasification of Lignin at Different Reaction Conditions.
wood biomass and synthetic polymers mixtures. Part III: Character- Energy Fuels 2014, 28, 6371−6379.
isation of heavy products. J. Anal. Appl. Pyrolysis 2003, 67, 325−340. (59) Asmadi, M.; Kawamoto, H.; Saka, S. Gas- and solid/liquid-
(39) Zhou, L.; Wang, Y.; Huang, Q.; Cai, J. Thermogravimetric phase reactions during pyrolysis of softwood and hardwood lignins. J.
characteristics and kinetic of plastic and biomass blends co-pyrolysis. Anal. Appl. Pyrolysis 2011, 92, 417−425.
Fuel Process. Technol. 2006, 87, 963−969. (60) Sophonrat, N.; Sandström, L.; Johansson, A.-C.; Yang, W. Co-
(40) Shen, D.; Xiao, R.; Gu, S.; Luo, K. The pyrolytic behavior of pyrolysis of Mixed Plastics and Cellulose: An Interaction Study by Py-
cellulose in lignocellulosic biomass: A review. RSC Adv. 2011, 1, GC×GC/MS. Energy Fuels 2017, 31, 11078−11090.
1641−1660.
(41) Zhang, H.; Carlson, T. R.; Xiao, R.; Huber, G. W. Catalytic fast
pyrolysis of wood and alcohol mixtures in a fluidized bed reactor.
Green Chem. 2012, 14, 98−110.
(42) Zhang, X.; Li, J.; Yang, W.; Blasiak, W. Formation Mechanism
of Levoglucosan and Formaldehyde during Cellulose Pyrolysis. Energy
Fuels 2011, 25, 3739−3746.
(43) Dorado, C.; Mullen, C. A.; Boateng, A. A. H-ZSM5 Catalyzed
Co-Pyrolysis of Biomass and Plastics. ACS Sustainable Chem. Eng.
2014, 2, 301−311.
(44) Lee, K.-H. Composition of aromatic products in the catalytic
degradation of the mixture of waste polystyrene and high-density
polyethylene using spent FCC catalyst. Polym. Degrad. Stab. 2008, 93,
1284−1289.
(45) Jae, J.; Tompsett, G. A.; Foster, A. J.; Hammond, K. D.;
Auerbach, S. M.; Lobo, R. F.; Huber, G. W. Investigation into the
shape selectivity of zeolite catalysts for biomass conversion. J. Catal.
2011, 279, 257−268.
(46) Cheng, Y.-T.; Huber, G. W. Chemistry of Furan Conversion
into Aromatics and Olefins over HZSM-5: A Model Biomass
Conversion Reaction. ACS Catal. 2011, 1, 611−628.
(47) Liu, Q.; Wang, S.; Zheng, Y.; Luo, Z.; Cen, K. Mechanism
study of wood lignin pyrolysis by using TG−FTIR analysis. J. Anal.
Appl. Pyrolysis 2008, 82, 170−177.
(48) Hemberger, P.; Custodis, V. B. F.; Bodi, A.; Gerber, T.; van
Bokhoven, J. A. Understanding the mechanism of catalytic fast
pyrolysis by unveiling reactive intermediates in heterogeneous
catalysis. Nat. Commun. 2017, DOI: 10.1038/ncomms15946.
(49) Wang, K.; Kim, K. H.; Brown, R. C. Catalytic pyrolysis of
individual components of lignocellulosic biomass. Green Chem. 2014,
16, 727−735.
(50) Zhang, Z.; Hirose, T.; Nishio, S.; Morioka, Y.; Azuma, N.;
Ueno, A.; Ohkita, H.; Okada, M. Chemical Recycling of Waste
Polystyrene into Styrene over Solid Acids and Bases. Ind. Eng. Chem.
Res. 1995, 34, 4514−4519.

473 DOI: 10.1021/acs.energyfuels.9b03135


Energy Fuels 2020, 34, 460−473

You might also like