Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
27 views52 pages

Final Report Draft

This project report details the simulation and optimization of hydrogen production from waste plastics, specifically polypropylene, high-density polyethylene, polystyrene, and polyethylene terephthalate, using gasification. The study evaluates the effects of various operating parameters on product yields, demonstrating that hydrogen and carbon monoxide yields are maximized at moderate air flow rates and elevated temperatures. The findings support the feasibility of converting plastic waste into hydrogen-rich syngas, contributing to sustainable waste management and clean energy solutions.

Uploaded by

macleodben205
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
27 views52 pages

Final Report Draft

This project report details the simulation and optimization of hydrogen production from waste plastics, specifically polypropylene, high-density polyethylene, polystyrene, and polyethylene terephthalate, using gasification. The study evaluates the effects of various operating parameters on product yields, demonstrating that hydrogen and carbon monoxide yields are maximized at moderate air flow rates and elevated temperatures. The findings support the feasibility of converting plastic waste into hydrogen-rich syngas, contributing to sustainable waste management and clean energy solutions.

Uploaded by

macleodben205
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 52

PROJECT REPORT

ON

“Simulation And Optimization of Production of Hydrogen from


Waste Plastics”

UNDER THE
GUIDANCE OF

Dr. Naresh Thota

DEPARTMENT OF CHEMICAL ENGINEERING


NATIONAL INSTITUTE OF TECHNOLOGY, WARANGAL
April, 2025

SUBMITTED BY: -

Sowthesh V(21CHB0A50)

Dhanumjay Atukuri(21CHB0B07) Yasir Parvez(21CHB0A58)


Approval Sheet
This Project Work entitled “Simulation and Optimization of
Production of Hydrogen from Waste Plastics”
by “Sowthesh V(21CHB0A50), Dhanumjay Atukuri(21CHB0B07), Yasir
Parvez(21CHB0A58)” is approved for the degree of B. Tech in Chemical
Engineering.

Examiners

_______________________

_______________________

________________________

_______________________

_______________________

Supervisor

______________________

Chairman

______________________

Date: 16-04-2025 Place: ________________


Declaration

I declare that this written submission represents my ideas in my own words and
where others’ ideas or words have been included, I have adequately cited and
referenced the original sources. I also declare the I have adhered to all principles
of academic honesty and integrity and have not misrepresented or fabricated or
falsified any idea/data/fact/source in my submission. I understand that any
violation of the above will cause for disciplinary action by the Institute and can
also evoke penal action from the sources which have thus not been properly
cited or from whom proper permission has not been taken when needed.

Roll No. Name of the student Signature of the student

21CHB0A50 Sowthesh V

21CHB0B07 Dhanumjay Atukuri

21CHB0A58 Yasir Parvez

Date:16-04-2025
Certificate

This is to certify that the Project work entitled


“Simulation And Optimization of Production of Hydrogen from
Waste Plastics”
is a bonafide record of work carried out by “Sowthesh V(21CHB0A50),
Dhanumjay Atukuri(21CHB0B07), Yasir Parvez (21CHB0A58)” submitted to
the faculty of Department of Chemical Engineering, in partial fulfilment of
the requirements for the award of the degree of Bachelor of Technology in
Chemical Engineering at National Institute of Technology, Warangal during
the academic year 2024-25.

Prof. Vidya Sagar Dr. Naresh Thota,


Head of the Department Assistant Professor,
Department of Chemical Department of Chemical
Engineering, Engineering,
NIT Warangal. NIT Warangal.
Contents
1. ABSTRACT.......................................................................................................................1
2. INTRODUCTION.............................................................................................................2
2.1. HYDROGEN:..................................................................................................................2
2.2. TYPES OF HYDROGEN...................................................................................................4
2.3. CHALLENGES WITH HYDROGEN:..................................................................................7
2.4. PLASTIC WASTES (PP, PS, HDPE, PET)......................................................................7
3. LITERATURE REVIEW...............................................................................................10
3.1. PROXIMATE AND ULTIMATE ANALYSIS.......................................................................10
3.2. DRYING OF NON-CONVENTIONAL SOLIDS:.................................................................11
3.3. RYIELD REACTOR:......................................................................................................12
3.4. GASIFICATION OF PLASTICS:.......................................................................................13
3.5. EFFECTS OF INERT ON GASIFICATION:........................................................................16
4. RESEARCH OBJECTIVES...........................................................................................18
5. METHODOLOGY..........................................................................................................19
5.1. INTRODUCTION...........................................................................................................19
5.2. COMPONENTS USED....................................................................................................19
5.2.1. Nonconventional Components................................................................................19
5.2.2. Conventional Components......................................................................................20
5.2.3. Characteristics of plastics.......................................................................................21
5.3. PROCESS MODEL DEVELOPMENT................................................................................21
5.3.1. Physical property method........................................................................................21
5.3.2. Model Assumptions.................................................................................................22
5.3.3. Model Flowsheet.....................................................................................................22
5.3.4. Model Description...................................................................................................22
6. RESULTS AND DISCUSSION......................................................................................27
6.1. EFFECT OF INPUT AIR FLOW RATE ON OUTLET COMPOSITION....................................27
6.2. EFFECT OF GASIFIER TEMPERATURE ON THE OUTLET COMPOSITION..........................31
6.3. EFFECT OF GASIFIER PRESSURE ON OUTLET COMPOSITION.......................................35
6.4. EFFECT OF STEAM ON GASIFICATION PRODUCTS.......................................................38
7. CONCLUSION................................................................................................................40
8. FUTURE SCOPE............................................................................................................42
9. REFERENCES................................................................................................................43
List of Figures
Figure 1 Global Demand of Hydrogen by Sector.......................................................................3
Figure 2 Types of Hydrogen.......................................................................................................4
Figure 3 Global Primary Waste generation by Polymer.............................................................7
Figure 4 Components Used in the Simulation..........................................................................21
Figure 5 Flowsheet of the Process............................................................................................22
Figure 6 Stoichiometry for the Drying of HDPE......................................................................23
Figure 7 Mole Fraction of Compounds Leaving the RYield reactor........................................24
Figure 8 Mole Fraction of Components leaving the Pyrolysis reactor.....................................25
Figure 9 Flowrate of H2 in the outlet by varying the inlet air flow rate....................................27
Figure 10 Flowrate of CO2 in the outlet by varying the inlet air flow rate...............................28
Figure 11 Flowrate of CO in the outlet by varying the inlet air flow rate................................29
Figure 12 Flowrate of CH4 in the outlet by varying the inlet air flow rate...............................30
Figure 13 Flowrate of H2 in the outlet by varying Gasifier Temperature.................................31
Figure 14 Flowrate of CO in the outlet by varying Gasifier Temperature................................32
Figure 15 Flowrate of CO2 in the outlet by varying Gasifier Temperature...............................33
Figure 16 Flowrate of CH4 in the outlet by varying Gasifier Temperature...............................34
Figure 17 Flowrate of H2 in the outlet by varying Gasifier Pressure........................................35
Figure 18 Flowrate of CO in the outlet by varying Gasifier Pressure......................................35
Figure 19 Flowrate of CO2 in the outlet by varying Gasifier Pressure.....................................36
Figure 20 Flowrate of CH4 in the outlet by varying Gasifier Pressure.....................................37
Figure 21 Effect of H2 flow on Introduction of steam...............................................................38
Figure 22 Modified Flowsheet for an optimised approach of production of hydrogen by
routing the moisture to the gasifier....................................................................................39
1. ABSTRACT

The growing accumulation of plastic waste poses significant environmental challenges, while
the global demand for clean hydrogen as a sustainable energy carrier continues to rise. This
project investigates the thermochemical conversion of various waste plastics—polypropylene
(PP), high-density polyethylene (HDPE), polystyrene (PS), and polyethylene terephthalate
(PET)—into hydrogen-rich syngas through gasification, using Aspen Plus process simulation
software. The study aims to evaluate the influence of key operating parameters, including air
flow rate, gasifier temperature, and pressure, on the yields of major gaseous products (H₂,
CO, CO₂, and CH₄) for each plastic type.
The process flowsheet was developed to simulate the drying, pyrolysis, and gasification of
plastic feedstocks, with a focus on equilibrium-based modeling in the absence of dedicated
oxidation and reduction reactors. The simulation workflow involved decomposing non-
conventional plastic components into their elemental constituents, followed by high-
temperature gasification in an RGibbs reactor. Parametric studies were conducted by
systematically varying air flow, temperature, and pressure to elucidate their effects on product
distribution.
Results indicate that hydrogen and carbon monoxide yields are maximized at moderate air
flow rates and elevated temperatures (800–1200°C), with PP and HDPE consistently
outperforming PS and PET due to their higher hydrogen content and hydrocarbon structure.
The CO production exhibited a characteristic bell-shaped curve with respect to air flow,
peaking at intermediate values before declining as complete oxidation to CO₂ became
dominant. Methane production was highest at lower temperatures and pressures, particularly
for PP and HDPE, but decreased sharply with increasing temperature due to enhanced steam
reforming and thermal cracking. Conversely, CO₂ and CH₄ yields increased with gasifier
pressure, while H₂ and CO yields declined, in accordance with Le Chatelier’s principle and
the thermodynamics of gasification reactions.
The simulation results validate the feasibility of converting waste plastics into hydrogen-rich
syngas via gasification and highlight the critical role of process conditions and feedstock
composition in optimizing product yields. This study provides valuable insights for the design
and operation of waste-to-hydrogen systems, supporting the development of sustainable
solutions for plastic waste management and clean energy production.
2. INTRODUCTION

Hydrogen energy has emerged as a promising alternate for addressing the pressing challenges
of climate change and transitioning towards a sustainable energy future. With its potential to
decarbonize hard-to-abate sectors and store excess renewable energy, hydrogen presents a
versatile solution to mitigate greenhouse gas emissions and enhance energy security.

2.1. Hydrogen

Hydrogen is the lightest and simplest element available in the universe. Under normal
conditions it exists as colourless, odourless, tasteless gas made up of two hydrogen
atoms. Even though it is abundant it is usually found combined with other elements like
in water. Hydrogen is highly flammable, when it burns in the presence of oxygen it
produces a lot of heat and one and only product which is water which makes it a clean
fuel. It is non-toxic, it has low density, because of this it must be stored carefully under
high pressure at low temperatures. Hydrogen has an awesome energy storage capacity
and the energy contained in 1 kg of hydrogen is about 120 MJ (=33.33 kWh), which
exceeds double of most conventional fuels. Since it is light weight, a large volume is
needed to store the same amount of energy as other fuels.
One of its most prominent applications is in transportation, were hydrogen fuel cells
power vehicles. In a fuel cell, hydrogen reacts with oxygen from the air to produce
electricity, powering electric motors and emitting only water vapor as a byproduct. This
process offers a clean and efficient alternative to traditional internal combustion
engines. Its major industrial application is in petroleum refining, where hydrogen is
used to remove impurities from crude oil and convert heavy hydrocarbons into lighter,
more valuable products like gasoline and diesel.
Figure 1 Global Demand of Hydrogen by Sector

This chart tracks how global hydrogen demand has changed over time and how it's
expected to evolve through 2050. From 1980 to 2018, the demand for hydrogen grew
steadily—from 25 million tonnes in 1980 to 74 million tonnes by 2018. During those
years, hydrogen was mainly used in refining and producing industrial ammonia, and the
growth was relatively slow and stable, averaging around 3% per year. The data up until
2018 is based on historical figures, while everything beyond that is a projection.
Looking ahead, there's a major shift on the horizon. Between 2030 and 2050, hydrogen
demand is expected to skyrocket, nearly tripling to 287 million tonnes. This boom is
largely driven by clean energy initiatives and the push to decarbonize various industries.
Sectors like transport, shipping, construction, and steel are projected to become major
consumers of hydrogen, with transport alone growing at over 20% per year.
Interestingly, the need for hydrogen in traditional refining processes is expected to
decline. Overall, the forecast reflects a future where hydrogen plays a central role in
building a more sustainable, low-carbon world.
Figure 2 Types of Hydrogen

2.2. Types of Hydrogen

Red, Pink, and Purple Hydrogen

These types of hydrogen are all produced using the electrolysis of water, but the
electricity used comes from nuclear energy. The terms red, pink, and purple are often
used interchangeably depending on the country or source. In this process, electricity
from a nuclear power plant is used to split water (H₂O) into hydrogen (H₂) and oxygen
(O₂) using an electrolyser. Since nuclear power does not emit carbon dioxide, this
hydrogen is considered low-emission and clean. It’s especially useful in countries that
already have a strong nuclear energy infrastructure.
While red, pink, and purple hydrogen all refer to hydrogen produced by electrolysis
powered by nuclear energy, the names vary mainly due to regional preferences or
emphasis. In some countries or research circles, pink hydrogen is the most commonly
used term. Others might use red hydrogen to specifically refer to hydrogen from older
or traditional nuclear reactors, while purple might be used when referring to advanced
nuclear technologies or nuclear thermal methods. However, the concept is the same
using low-carbon electricity from nuclear power to split water into hydrogen and
oxygen.

Blue Hydrogen

Blue hydrogen is produced by a method called steam methane reforming (SMR), where
natural gas (mostly methane) is heated with steam to produce hydrogen and carbon
dioxide (CO₂). What makes it "blue" is that the CO₂ emissions are captured and stored
underground using carbon capture and storage (CCS) technology. This reduces its
environmental impact. Blue hydrogen is seen as a cleaner alternative to grey hydrogen
and is considered a transitional solution toward greener energy systems.

Turquoise Hydrogen

Turquoise hydrogen is created through a process called methane pyrolysis, where


methane (CH₄) is heated in the absence of oxygen to split it into hydrogen gas and solid
carbon. This process avoids releasing carbon dioxide, making it cleaner than traditional
fossil-fuel-based hydrogen. However, the environmental benefit depends on whether the
heat used in the process comes from renewable energy sources. Since it produces solid
carbon, which is easier to handle than gaseous CO₂, it has promising potential but is
still under development.

Gray Hydrogen

Gray hydrogen is also made using steam methane reforming (SMR), like blue hydrogen,
but without capturing the carbon dioxide. This means that all CO₂ is released into the
atmosphere, making it the most commonly used but most polluting form of hydrogen
today. It is inexpensive to produce and widely used in industries such as oil refining and
fertilizer production, but it has a significant environmental impact due to its high carbon
emissions.

Black and Brown Hydrogen

Black and brown hydrogen are both produced by gasifying coal — bituminous coal in
the case of black hydrogen and lignite (brown coal) for brown hydrogen. The coal is
heated in the presence of oxygen and steam to break down its components, releasing
hydrogen along with a large amount of CO₂ and other pollutants. These methods are the
dirtiest and most carbon-intensive because coal is one of the most polluting fossil fuels.
Some countries still use these methods due to their heavy dependence on coal.
Black and brown hydrogen are both produced through the gasification of coal, which
involves heating coal with steam and limited oxygen to release hydrogen gas along with
large amounts of carbon dioxide (CO₂) and other pollutants. The key difference lies in
the type of coal used. Black hydrogen comes from bituminous (hard) coal, while brown
hydrogen comes from lignite (soft or brown) coal. Lignite contains more moisture and
less energy, making brown hydrogen even more polluting than black hydrogen. Both are
considered highly carbon-intensive and are the least environmentally friendly forms of
hydrogen.

Gold Hydrogen

Gold hydrogen refers to naturally occurring hydrogen found underground that can be
extracted in an economically viable and environmentally friendly way. This is a new
and experimental approach. If successful, it could allow hydrogen to be obtained
directly from the Earth without the need for complex processing or harmful emissions.
Gold hydrogen is still under study, but it shows promise as a natural and clean hydrogen
source if extraction methods can be developed responsibly.

White Hydrogen

White hydrogen is naturally present in the Earth's crust and sometimes escapes through
cracks in rocks. It forms through natural geological processes such as the reaction of
water with certain types of minerals (like iron-rich rocks). Unlike other types, this
hydrogen doesn’t need to be manufactured — it's already there. However, it is rarely
tapped into and is still being researched to determine how it can be safely and efficiently
extracted for large-scale use.

Green Hydrogen

Green hydrogen is the cleanest and most environmentally friendly type. It is produced
using a process called electrolysis, where electricity splits water (H₂O) into hydrogen
and oxygen. What makes it "green" is that the electricity comes from renewable sources
such as solar, wind, or hydropower. This means the entire process is carbon-free,
making green hydrogen ideal for achieving a low-carbon or net-zero future. Though
currently more expensive than other types, costs are expected to drop as renewable
energy becomes more common.

Yellow Hydrogen

Yellow hydrogen is also made through electrolysis, but the electricity comes from the
general power grid rather than strictly from renewables. This grid electricity could be a
mix of coal, gas, nuclear, or renewables, depending on the region. As a result, the
cleanliness of yellow hydrogen varies, depending on how clean the local grid is. It’s a
flexible but less consistent method in terms of environmental impact.

2.3. Challenges with Hydrogen:

The challenges faced by hydrogen fuel is primarily related to production, storage,


distribution, and infrastructure. While hydrogen is abundant, its production often relies
on energy-intensive processes such as steam methane reforming, which can emit
greenhouse gases unless renewable energy sources are utilized. Additionally, hydrogen
has low energy density by volume, requiring large storage tanks or compression
systems, which can be costly and bulky, limiting its practicality for transportation. The
need to maintain high pressure and low temperature for the storage of hydrogen is
practically difficult. Safety concerns also arise due to hydrogen's flammability and the
need for specialized handling protocols. By overcoming these challenges, hydrogen can
serve as a viable and sustainable fuel for future.

2.4. Plastic Wastes (PP, PS, HDPE, PET)

Plastic management is a major environmental challenge due to its non-biodegradable


nature. Mainly, four frequently trashed plastics are Polypropylene (PP), Polystyrene
(PS), Polyethylene Terephthalate (PET), High Density Polyethylene (HDPE).

Figure 3 Global Primary Waste generation by Polymer


This chart "Global primary plastic waste generation by polymer, 2019" tells us about the
total plastic waste generated globally by different types of polymers. It offers insights
into both the quantity of waste each polymer produces and its recyclability, which is
visually represented through a color-coded system. This data originates from the OECD
(2022) and is curated by Our World in Data.
At the top of the list is Polypropylene (PP), generating 62 million tonnes of waste in
2019. It is color-coded in blue, indicating that it is only moderately recycled, despite its
massive contribution to plastic waste. This is concerning because it shows that the most
used plastic is not being effectively reintegrated into the production cycle. LDPE and
LLDPE (Low-Density Polyethylene and Linear Low-Density Polyethylene) follow
closely with 49.21 million tonnes, also in blue, and HDPE (High-Density Polyethylene)
comes in at 44.71 million tonnes, coloured green, indicating it is widely recycled. This
makes HDPE one of the more sustainable options among the top contributors.
Another significant category is “Other”, accounting for 57.56 million tonnes, and
Fibres, contributing 51.76 million tonnes. Both are marked in violet, which represents
unknown recyclability. The presence of such a large portion of plastic waste under
ambiguous recycling status highlights a critical gap in data and waste management
infrastructure. These categories likely include mixed or composite plastics used in
textiles, consumer goods, and industrial applications, which are more difficult to
separate and recycle.
PET (Polyethylene Terephthalate), used widely in beverage bottles and food containers,
contributes 24.81 million tonnes and is green-coded, meaning it is one of the most
commonly recycled plastics globally. This reflects successful recycling campaigns and
infrastructure in place for PET. In contrast, PVC (Polyvinyl Chloride) at 21.2 million
tonnes, Polystyrene (PS) at 15.17 million tonnes, and Polyurethane (PUR) at 11.34
million tonnes are all marked in red, indicating that they are usually non-recycled.
These materials often pose significant recycling challenges due to their chemical
composition, contamination, or the release of harmful substances during processing.
Less common plastic waste generators like ABS, ASA, and SAN (used in durable goods
like electronics and vehicles) generate 7.28 million tonnes, while elastomers (used in
tyres) account for 5.09 million tonnes. Both are marked in violet, reflecting a lack of
clarity about their recyclability. Bioplastics, often considered a sustainable alternative,
generated 2.1 million tonnes and are marked in orange, meaning they have limited
recyclability. This is important because while marketed as “eco-friendly,” bioplastics
often require specific industrial conditions to degrade and are not always compatible
with standard recycling systems.
At the lower end of the scale, road marking coatings and marine coatings contribute
579,552 tonnes and 483,028 tonnes, respectively. Although their quantities are smaller,
these substances are also marked in violet, and their environmental impact may be
disproportionately large due to their persistence and spread in marine and urban
environments.
The color-coding system used in the chart is essential for understanding the recyclability
of each polymer type:
 Green denotes polymers that are widely recycled (e.g., PET, HDPE),
 Blue indicates those that are moderately recycled (e.g., PP, LDPE),
 Orange suggests limited recyclability (e.g., Bioplastics),
 Red shows usually non-recycled materials (e.g., PVC, PS),
 Violet implies unknown recyclability, typically due to complex or mixed-material
compositions.
PP is highly popular in textile, automotive components, and packaging because of its
heat-resistance and toughness. It has high hydrogen content making it an ideal
component for hydrogen production via method named gasification. Due to its
lightweight, PS is commonly found in disposable cups, insulation, and packaging but it
generates high carbon monoxide levels when gasification is done. HDPE is found in
bottles, containers, and pipes, these are highly recyclable and has an excellent hydrogen
yield potential. PET is mainly used in beverage bottles and food packaging, has high
oxygen content, which results in lower hydrogen production efficiency.
Recycling and waste-to-energy methods such as gasification can turn these plastics into
useful fuels, reducing the landfill waste and environmental pollution.
3. LITERATURE REVIEW

3.1. Proximate and Ultimate Analysis

Proximate analysis determines the physical and thermal properties of plastic waste by
measuring four key parameters:
 Fixed Carbon (FC): The solid carbonaceous residue left after volatile components
are drive off. It contributes to char formation during gasification.
 Volatile Matter (VM): The fraction of material that vaporizes when heated,
producing combustible gases such as hydrogen (H2), carbon monoxide (CO), and
methane (CH4). Higher volatile matter improves gasification efficiency.
 Ash Content: The inorganic residue remaining after complete combustion. Ash
affects gasification performance as it does not contribute to energy production and
may cause slagging in reactors.
 Moisture Content (MC): The water present in plastic waste, which affects thermal
efficiency. Moisture must be minimized as it consumes energy during evaporation.
Ultimate analysis measures the elemental composition of plastics, which influences
their combustion behaviour. The primary elements analysed include:
 Carbon (C): The main contributor to syngas production, influencing CO and CO 2
levels. Higher carbon content enhances energy yield.
 Hydrogen (H): Plays a crucial role in hydrogen-rich syngas production. Plastics
like PP and HDPE have high hydrogen content, making them suitable for hydrogen
production.
 Oxygen (O): Impacts combustion and gasification efficiency. Higher oxygen
content (as seen in PET) leads to lower oxygen yield.
 Nitrogen (N) and Sulphur (S): Present in trace amounts; their combustion can
form pollutants like NOx and SO2.
 Chlorine (Cl): Found mainly in PVC, forming harmful hydrogen chloride (HCl)
emissions during thermal decomposition.

Table 1 Proximate and Ultimate Analysis of Various Plastics

Polypropylene High Density Polystyrene Polyethylene


(PP) Polyethylene (PS) Terephthalate
Yao D et al. (HDPE) Encinar JM (PET)
Yao D et al. et al. Encinar JM et
al.
Proximate Analysis
(mass fraction %)

Fixed Carbon 0.03 0 0 13.15

Volatile Matter 99.97 94.77 99.8 86.85

Ash 0 4.98 0.2 0

Moisture Content (MC) 0 0.25 0 0

Ultimate Analysis
(mass fraction %)

Carbon 84.14 78.18 90.02 62.5

Hydrogen 14.96 12.84 8.48 4.21

Nitrogen 0.23 0.06 0 0

Sulphur 0.24 0.08 0 0

Chlorine 0 0 0 0

Oxygen 0.43 3.61 1.3 33.29

3.2. Drying of Non-conventional Solids:

Drying non-conventional solids in Aspen Plus involves defining components as


nonconventional solids, specifying their physical properties, and selecting appropriate
thermodynamic models. Nonconventional solids, such as coal or biomass, are treated as
heterogeneous mixtures requiring specialized property methods like HCOALGEN for
enthalpy and DCOALIGT for density.
Practical implementation involves configuring dryer blocks (e.g., DRYER, RSTOIC)
and validating against experimental data. For instance, the Aspen tech dryer model uses
a stoichiometric reactor (RSTOIC) to simulate moisture removal, followed by a flash
separator (FLASH2) to split dried solids and humid gas. Sensitivity analyses on
variables like inlet air flow and initial moisture content help optimize energy use, with
excess airflow reducing exit temperatures to prevent auto-ignition risks in biomass
storage. Challenges include accurately representing particle size distributions and
selecting heat capacity models, where cubic interpolation methods better approximate
biomass behavior compared to Kirov’s equation.
Advanced configurations integrate drying with adsorption or heat recovery to enhance
efficiency. Synergistic processes combining adsorbent dehumidification with drying can
reduce energy consumption by up to 70% for low-temperature applications. For
example, microporous adsorbents regenerate moisture-laden air streams, while
mesoporous materials recover heat from exhaust gases. These optimizations are
modelled using Aspen’s Design-Spec or Calculator tools to dynamically adjust
operating parameters. References to studies like Chen et al. (2012) on rice straw
drying and Atuonwu et al. (2013) on adsorbent-enhanced drying further validate these
approaches, highlighting Aspen Plus flexibility in scaling laboratory data to industrial
processes [1].

3.3. RYield Reactor:

The RYield reactor in Aspen Plus is a stoichiometric reactor module used to model
processes where the exact reaction kinetics are unknown, but the yield distribution of
products can be specified based on empirical data or thermodynamic decomposition. It
is particularly useful for simulating the decomposition of non-conventional components
(e.g., plastics, biomass, or coal) into their elemental constituents or simpler compounds.
Non-conventional components (e.g., plastics) are defined using
HCOALGEN and DCOALIGT models in Aspen Plus to specify their elemental
composition. Products are conventional components (e.g., C, H₂, CO, CH₄) whose
yields are calculated from the feedstock’s ultimate/proximate analysis. The reactor
operates under user-defined temperature and pressure conditions, often linked to
downstream processes like gasification or combustion.
Han et al., 2023 RYield modelled the pyrolysis of polypropylene (PP) and polystyrene
(PS) into syngas precursors. The yields were derived from experimental data to ensure
alignment with real-world gasification behaviour. Nikoo & Mahinpey, 2008 RYield
decomposed biomass into volatile gases (CO, H₂, CH₄) and char. The yields were based
on the feedstock’s proximate analysis (moisture, volatiles, fixed carbon), enabling
integration with downstream gasification reactions. Aspen Tech, 2023 Aspen Tech
recommends using RYield for coal devolatilization by specifying char and volatiles
yields. This approach simplifies complex pyrolysis kinetics into a yield-driven model.
Mittal et al., 2006 In tire gasification, RYield decomposed tires into carbon, sulphur,
and volatiles. The model used ultimate analysis (e.g., 85.65% C, 1.43% S) to predict
syngas composition in downstream reactors.

3.4. Gasification of Plastics:

Gasification of plastics in Aspen Plus involves modeling complex thermochemical


processes to convert waste polymers into syngas, with studies focusing on reactor
design, reaction mechanisms, and parameter optimization. For instance, Kannan et
al. developed a fluidized bed gasification model for polyethylene (PE) using Aspen
Plus, incorporating drying, pyrolysis, combustion, and gasification stages. Their
sensitivity analysis revealed that an equivalence ratio (ER) of 0.25 maximizes syngas
yield, while higher steam-to-fuel ratios (>0.5) reduce thermal efficiency due to
excessive heat absorption[2]. The model assumed steady-state conditions and
decomposed PE into volatile components and char, with gasification occurring in both
bubbling bed (RCSTR) and freeboard (RPFR) reactors [5].
The reaction chemistry in plastic gasification involves pyrolysis-driven devolatilization
followed by char reduction. Saebea et al. highlighted that polyethylene and
polypropylene decompose into light hydrocarbons (e.g., CH₄, C₂H₄) and hydrogen
during pyrolysis, which later reform into syngas (H₂, CO) via steam and CO₂
interactions in RGibbs reactors[3]. Aspen Plus modules like RYield and RGibbs are
commonly used to simulate these stages, with non-conventional plastic components
decomposed into elemental constituents (C, H, O) based on ultimate analysis[7]. For
example, PE’s high carbon content (83.3 wt%) and low ash (0.1 wt%) make it ideal for
syngas production compared to biomass blends.
Key process parameters include equivalence ratio, steam-to-carbon (S/C) ratio, and
temperature. Increasing ER from 0.1 to 0.5 elevates gasification temperatures from
565°C to 1,734°C but reduces H₂ content beyond ER=0.25 due to excessive
oxidation. Steam injection enhances H₂ production via water-gas shift reactions, with
optimal S/C ratios of 1.2–1.4 ensuring complete carbon conversion[7]. However,
excessive steam (>0.5 ratio) dilutes syngas and lowers efficiency. Bed height and
reactor volume also influence conversion rates, with taller beds reducing feedstock
conversion due to uneven fluidization.
Co-gasification of plastics with biomass, such as rice husk (RH), improves process
sustainability. A study blending HDPE with RH found that increasing HDPE content
from 0% to 100% raised syngas yield from 0.60 kg/kg to 2.2 kg/kg of feed, while net
plant efficiency improved from 35% to 68% due to HDPE’s higher carbon
content[6]. CO₂ capture and utilization (CCU) further enhanced sustainability, with
90% of captured CO₂ recycled into reforming reactions for HDPE-rich
feeds[6]. However, biomass’s high ash content (e.g., 19.7% in RH) necessitated
additional cleanup steps, increasing energy demands[6].
Tar and char management are critical challenges. Toledo et al. noted that polyolefins
like PE produce minimal tar under high temperatures (>1,250°C) and prolonged
residence times, but aromatic plastics (e.g., PET, PS) generate persistent tars requiring
catalytic cracking[4]. Wang et al. observed that CO₂-assisted gasification of PET
increased char porosity, enhancing Boudouard reactions (C + CO₂ → 2CO) and
reducing solid residues[4]. Aspen Plus simulations often exclude tar formation by
assuming ideal gasification conditions, though this may overestimate syngas purity[7].
Environmental and efficiency analyses reveal trade-offs. Pure HDPE gasification
achieves 68% plant efficiency but relies on fossil-derived plastics, while RH blends
reduce carbon footprint at the cost of lower output[7]. Marchese et al. compared end-
products, finding methanol synthesis most viable due to moderate energy demands (2.2
kg CO₂-eq/kg PE) versus hydrogen production, which requires optimization to compete
with waste-to-energy incineration. Oxy-steam gasification models showed a maximum
hydrogen energy efficiency (HEE) of 36.95% at 400°C steam temperatures, with multi-
objective optimization identifying ER=0.25 and 25% plastic content as ideal.
The gasification of plastics involves a series of thermochemical reactions modelled in
Aspen Plus to optimize syngas production. Below are key reactions and their
mechanisms, derived from simulation studies and experimental validations:

Boudouard Reaction

 C+CO2→2CO (ΔH=+172 kJ/mol)


This reaction dominates in CO2-rich environments, enhancing CO production by
converting char and carbon dioxide. It is critical in CO 2-assisted gasification, where
pore development in char improves reactivity[5][10].

Char Gasification (Water-Gas Reaction)


 C+H2O→CO+H2 (ΔH=+131 kJ/mol)
Steam reacts with solid carbon (char) to produce hydrogen and carbon monoxide.
This reaction is temperature-sensitive, with higher steam-to-carbon ratios favoring
H2 yield[5][10].

Water-Gas Shift Reaction

 CO+H2O↔CO2+H2 (ΔH=−41.2 kJ/mol)


This equilibrium reaction adjusts the H2/CO ratio in syngas. It is exothermic and
favored at lower temperatures (300–500°C) [5][10].

Methanation

 C+2H2→CH4 (ΔH=−74.8 kJ/mol)


Occurs under high-pressure, hydrogen-rich conditions, producing methane. While
undesirable for syngas purity, it contributes to the heating value of the gas [5][10].

Steam Reforming of Methane

 CH4+H2O→CO+3H2 (ΔH=+206 kJ/mol)


Catalysed by Ni-based materials (e.g., Ni-Al2O3), this endothermic reaction
converts methane and steam into hydrogen and CO at temperatures >800°C[9].

Plastic Depolymerization and Reforming

 Polyethylene (PE) and polypropylene (PP) decompose into radicals (e.g., R˙, ˙OH)
at high temperatures (2500–3500 K), which react with steam:
 CnHm+H2O→CO+H2
 This step is critical for breaking long-chain hydrocarbons into syngas components,
with steam-to-plastic ratios >1.5 enhancing H2 yield by 2.5 [8].

Key Parameters Influencing Reactions

 Temperature: Higher temperatures (800–950°C) favour endothermic reactions


(Boudouard, steam reforming) and reduce tar[9][8].
 Air-to-Plastic Ratio: Optimal A/P ratios maximize H2 yield without excessive heat
loss[9][8].
 Catalysts: Ni-Al2O3 enhances reforming efficiency and reduces tar by 60–80%[9]
[10].
3.5. Effects of inert on Gasification:

Inert materials play a critical role in gasification processes modeled via Aspen Plus,
primarily influencing temperature regulation, reaction kinetics, and syngas composition.
In bubbling fluidized bed gasifiers, inert bed materials like olivine or quartz sand are
heated during char combustion, stabilizing reactor temperatures and enhancing heat
transfer to endothermic gasification reactions[10]. Higher gasification temperatures
(800–1000°C) promoted by inert beds increase CO and H₂ yields but reduce CH₄
content and cold gas efficiency due to equilibrium shifts[10][11]. Inerts also affect
hydrodynamics: smaller particle sizes (0.2–0.5 mm) improve fluidization, reduce tar
formation, and enhance gas-solid contact, as demonstrated in simulations of Prosopis
Juliflora gasification[10][9].
The inert content within biomass feedstocks, such as ash, directly impacts syngas
quality. Aspen Plus models define ash as a non-conventional component using
proximate and ultimate analyses (PROXANAL/ULTANAL), which influence enthalpy
calculations and mass balances[3]. High ash content reduces effective carbon available
for gasification, lowering syngas heating value and necessitating higher equivalence
ratios (ER) to maintain combustion stability[5]. For instance, increasing the air-to-
biomass ratio from 0.2 to 1.2 in a multistage gasifier improved cold gas efficiency but
introduced more inert N₂ from air, diluting syngas[10][11].
Inerts from gasifying agents like air (N₂) or steam further alter process efficiency. Air-
based gasification introduces significant N₂, diluting H₂ and CO concentrations and
reducing syngas lower heating value (LHV)[10]. Studies comparing air and oxygen-
enriched gasification show that minimizing N₂ content increases LHV from ~4 MJ/m³
to over 10 MJ/m³[11]. Conversely, steam as a gasifying agent avoids inert dilution but
requires careful temperature control to prevent tar formation. Catalytic inerts like
olivine or dolomite, though non-reactive, indirectly reduce tar (e.g., naphthalene,
toluene) via enhanced cracking reactions, as observed in plastic waste gasification[9].
Aspen Plus simulations incorporate inerts through property methods (e.g., PR-BM) and
reactor configurations. Gibbs equilibrium models simplify inert effects by assuming
ideal mixing, while kinetic models (CSTR) require explicit reaction rates and residence
times[11]. For example, fluidized bed hydrodynamics are modeled using Fortran
subroutines to account for inert-driven heat transfer and particle interactions
[11]. Validation against experimental data, such as the Güssing plant, highlights the
need to adjust ER and bed material size to align simulated and actual syngas
compositions. These approaches ensure accurate predictions of inert impacts on
scalability and optimization.
4. RESEARCH OBJECTIVES

 To investigate the potential of plastic waste as a feedstock for synthesis gas (syngas)
production through thermochemical conversion.
 To simulate the gasification process using appropriate modeling techniques and evaluate
key performance metrics such as syngas composition and yield.
 To analyze the influence of operating parameters such as temperature, pressure and air-
to-feed (A/F) ratio on syngas generation, methane generation and carbondiocide
generation
 To determine the optimal process conditions for maximizing hydrogen and carbon
monoxide production from plastic waste.
 To assess the effect of different types of plastic (e.g., polypropylene and polystyrene)
and their combinations on the quality and quantity of syngas produced.
 To contribute to sustainable waste management strategies by exploring value-added
utilization of non-recyclable plastic waste.
5. METHODOLOGY

5.1. Introduction

The present study focuses on the conversion of plastic waste into synthesis gas (syngas)
through a thermochemical gasification process. The methodology involves process
modelling and simulation using Aspen Plus software to analyse the behaviour of the
system under varying operational conditions. The approach is grounded in equilibrium-
based simulation techniques, wherein the chemical composition of the gas products is
estimated based on the minimization of Gibbs free energy.
Plastic waste, primarily polystyrene (PE), polypropylene (PP), high density
polyethylene (HDPE) and polyethylene terephthalate, is considered as the feedstock due
to their high carbon and hydrogen content. These polymers are initially decomposed
into their elemental constituents, which subsequently undergo multiple gas-phase
reactions in the presence of steam as a gasifying agent. The process parameters such as
temperature, pressure and air-to-feed (A/F) ratio are systematically varied to study their
influence on syngas yield and composition.
The simulation setup includes defining thermodynamic models, feedstock properties
(via ultimate and proximate analyses), and reaction conditions to closely replicate
practical gasification behaviour. The outcomes of the simulation provide insights into
the optimal conditions required to maximize hydrogen and carbon monoxide
production, and the overall energy value of the syngas. This methodology serves as a
foundation for evaluating the feasibility and performance of plastic waste gasification as
a sustainable waste-to-energy solution.

5.2. Components Used

In this study, a range of conventional and nonconventional components were defined


within the Aspen Plus simulation environment to accurately represent the
thermochemical conversion of plastic waste into synthesis gas (syngas).

5.2.1. Nonconventional Components

These represent complex materials that do not have a defined chemical formula,
typically used for solid waste inputs in simulation. Their composition is defined via
proximate and ultimate analyses.
PP (Polypropylene) – A hydrocarbon-based polymer used as a model plastic feedstock.
POLYS (Polystyrene) – A synthetic aromatic hydrocarbon polymer included as a
representative waste plastic.
PET (Polyethylene Terephthalate) – A common polyester plastic, modelled as a
nonconventional component.
HDPE (High-Density Polyethylene) – Another key plastic component in the feedstock
mix.
ASH – Represents the inert residue remaining after gasification, typically defined with
zero reactivity.

5.2.2. Conventional Components

These are standard chemical species with defined molecular formulas and
thermodynamic properties. They are primarily the products and intermediates of
gasification.
Table 2 Components used in the Simulation

Component Name Alias Function in Process

Carbon Monoxide CO Major syngas component

Carbon Dioxide CO₂ Byproduct from oxidation and reforming

Nitrogen N₂ Inert gas; used for drying

Oxygen O₂ Oxidizing agent for partial combustion

Hydrogen H₂ Valuable syngas product

Chlorine Cl₂ Possible contaminant from plastic

Ammonia NH₃ Formed from nitrogen in plastics

Sulfur S Impurity that can form SOx

Carbon (Graphite) C Residual carbon or char

Methane CH₄ Secondary fuel in syngas

Hydrogen Sulfide H₂S Formed from sulfur in impurity

Hydrogen Chloride HCl Forms from chlorinated plastics


Water H₂O Gasifying agent (steam) and product

Figure 4 Components Used in the Simulation

5.2.3. Characteristics of plastics

Four different types of waste plastics were chosen as fuels for hydrogen production. The
proximate analysis and ultimate analysis results of the samples were taken from the
literature as listed in Table 1

5.3. Process model development

5.3.1. Physical property method

Prior to running simulations, proper stream classes should be determined. Material


streams are classified by Aspen Plus into three types which are mixed, conventional
solids, and nonconventional solids. Aspen Plus libraries were utilized to identify the
thermodynamic properties of chemical components. Gases were specified as mixed sub-
streams, char as a conventional solid substream, plastics and ash as nonconventional
solid substreams. The plastics were specified as nonconventional components with their
ultimate and proximate analyses according to Table 1 due to their heterogeneous solid
structure. For calculating the enthalpy and density of the nonconventional solids, in-
built HCOALGEN and DCOALIGT algorithms were chosen. All physical
characteristics of the conventional components in the gasification process were
evaluated using the Peng-Robinson equation with Boston Mathias modification (PR-
MB), recommended for gasification process. In order to perform the modeling process,
the following assumptions are taken into consideration.

5.3.2. Model Assumptions

In the kinetic-free numerical model, both stoichiometric and non-stoichiometric


(Gibbs's free energy minimization method) thermodynamic equilibrium approaches
were applied. In this regard, the gasifier was modeled using the following assumptions.
 Steady state steady flow process and isothermal conditions are considered.
 All pressure drops in the system are neglected.
 All chemical reactions are in equilibrium and the residence time is long enough to
achieve chemical and thermodynamic equilibrium.
 Nitrogen, chlorine and sulfur in the feedstock if exist react to give ammonia (NH 3),
hydrogen chloride (HCl) and hydrogen sulfide (H2S).
 Ideal condition is considered for all gases.
 The ash is inert and does not involve in chemical reactions.
 The formation of tar and other heavy hydrocarbons are not considered in the model.
 All elements except ash participate in chemical reactions and contact one another
uniformly.

5.3.3. Model Flowsheet

Figure 5 Flowsheet of the Process


5.3.4. Model Description

The gasification process was mainly comprised of four sub-processes which are drying,
pyrolysis, reduction (gasification), and oxidation (combustion) and therefore these sub-
processes were represented by separate units in Aspen Plus simulator. First, the
PLASTIC stream is fed to the DRIER reactor to simulate the drying process. The SEP
DRY separator is used to extract the evaporated water (MOIST) from S2 stream. Then,
the dried feedstock (S4) enters the DECOMP reactor which is a RYield type reactor in
order to simulate the yield distribution after pyrolysis/decomposition of the
nonconventional feedstock with respect to the ultimate analysis of the plastics in Table1.
Drier
The dryer block is designed to remove moisture from the plastic waste feedstock before
it enters the gasification reactors. Even though plastics typically have low moisture
content, waste plastic can retain surface moisture or absorb water during collection,
washing, and storage.
For drying, inter gas, nitrogen is used. The source of nitrogen is air. Initially air is
heated to 150oC using heater (HEATER1), then oxygen and nitrogen are separated using
a separator (SEP1). The oxygen is used in the further process and the nitrogen is
separated and being used in the drier. The drier operates at 1bar pressure and 150 oC
temperature.
Aspen considers 1gm of non-conventional component as 1 mole of non-conventional.
So, in order to achieve the mass balance, the co-efficient of H2O is 0.0555084. This
represents that for 1 mole of plastic, 0.0555084 mole of H2O is formed. Also, the
conversion is set to 20% in the reaction.
PLASTIC → PLASTIC + 0.0555084 H2O
An example for HDPE drying is given below.
Figure 6 Stoichiometry for the Drying of HDPE

From the dryer, the moisture is removed by a separator block (FLASH) along with
nitrogen. The dried plastic is then sent to a decomposer reactor (Ryield).
Decomposer
In the gasification of plastic waste, one of the major modelling challenges arises from
the fact that plastic materials—such as polypropylene (PP), polyethylene (PE),
polystyrene (PS), and polyethylene terephthalate (PET)—are nonconventional
components in Aspen Plus. These materials do not have fixed chemical formulas and are
instead defined by their proximate and ultimate analyses. As such, they cannot directly
participate in chemical reactions in stoichiometric or equilibrium-based reactors without
first being decomposed into conventional components.
The breakdown of the plastic in the RYield reactor is performed according to the
ultimate analysis of each plastic type. The ultimate analysis provides the elemental
composition of the feedstock — typically the weight percentages of carbon (C),
hydrogen (H), oxygen (O), nitrogen (N), sulfur (S), and ash. The individual component
yields are mentioned in the reactor so as to simulate the breakdown of plastic polymers.
The reactor operates at 700oC and 1 bar pressure.
Figure 7 Mole Fraction of Compounds Leaving the RYield reactor

Ash is solid and does not participate in any reaction. In order to separate ash from the
remaining components, a cyclone separator is used. The ash then cooled and separated
out from the process from the stream ASHOUT. It separates the non-conventional
component present in the stream S5. Remaining components are sent to the pyrolysis
reactor.
Pyrolysis Reactor
The pyrolysis reactor is modeled as an RGibbs reactor in Aspen Plus. This reactor type
is used for systems where the precise reaction pathways and kinetics are either unknown
or complex, a typical case for the breakdown of plastics. Instead of defining specific
chemical reactions, the RGibbs reactor computes the chemical equilibrium by
minimizing the total Gibbs free energy of the system. Initially the block operates at
900.C and 5 bar pressure. The mole fraction of the outlet steam from the pyrolysis
reactor is given below.
Figure 8 Mole Fraction of Components leaving the Pyrolysis reactor

Separator SEP2:
SEP2 is a gas-solid separator used to remove char and inorganic components such as
HCl, NH3, and H2S from the product gas stream exiting the pyrolysis section. This
separator ensures that only the gaseous components proceed further in the process while
solid residues such as inorganic contaminants are removed as a separate stream. This
purification step is crucial to prevent downstream equipment fouling and to enhance the
quality of the product gas for subsequent cooling and purification stages.

Cooler COOLER1:
COOLER1 is introduced after SEP2 to lower the temperature of the cleaned gaseous
stream. The gases exiting the pyrolysis and separation stages are typically at high
temperatures, which are not suitable for further gas-liquid separation and compression
stages. This heat exchanger cools the gas to a temperature that is compatible with the
efficient operation of the subsequent separator (SEP3) and compressor, while also
potentially recovering heat for energy integration elsewhere in the process. The
temperature is reduced to atmospheric temperature of 25oC.
Separator SEP3:
SEP3 is a gas-liquid separator designed to remove condensed water from the cooled gas
stream. This step is essential to ensure that only dry gases enter the compressor and
pressure swing adsorption (PSA) unit, as the presence of moisture could interfere with
the adsorption process and reduce its effectiveness in separating hydrogen from other
gases.
Compressor COMP:
The COMP block represents a gas compressor that increases the pressure of the dry gas
stream to levels suitable for pressure swing adsorption. PSA processes typically operate
at elevated pressures (usually between 5–30 bar) to allow effective separation of
hydrogen based on differences in adsorption characteristics of the gas components. The
compressor is a key element in preparing the gas stream for hydrogen purification,
enabling higher hydrogen recovery and purity. In this process the compression is done
to a discharge pressure of 25 bars.
Separator SEP4:
SEP4 represents the pressure swing adsorption (PSA) unit in the simulation. This
separator is modeled to split the incoming compressed gas stream into two outputs: a
high-purity hydrogen stream and a tail gas or flue gas stream. The PSA unit exploits the
differences in adsorption capacity of the various gas components on specific adsorbent
materials, allowing hydrogen to pass through while other gases are temporarily
adsorbed and later desorbed during regeneration. This final step in the process ensures
the production of purified hydrogen suitable for downstream applications or storage.

6. RESULTS AND DISCUSSION

This section presents the key findings from the simulation of hydrogen production from
plastic waste using Aspen Plus. It highlights the behavior of various units involved in
the process, the effect of various parameters like temperature, pressure and flow rates of
the input. The results are analyzed based on parameters such as product composition
and hydrogen yield.

6.1. Effect of input air flow rate on Outlet composition

Effect on outlet H2 flow on Inlet air flow


70

60

50

40
H2 flow

ps
30 pp
hdpe
20 pet
10

0
0 500 1000 1500 2000 2500
Inlet air flow

Figure 9 Flowrate of H2 in the outlet by varying the inlet air flow rate

The graph illustrates hydrogen production (kg/hr) from different plastic feedstocks (PP,
HDPE, PS, and PET) as oxygen input increases, revealing distinct behaviours based on
polymer chemistry. Initially, PP demonstrates the highest hydrogen yield (~67 kg/hr)
followed by HDPE (~60 kg/hr), PS (~38 kg/hr), and PET (~19 kg/hr), directly
correlating with their hydrogen content and molecular structure. All feedstocks exhibit a
stable production plateau at low oxygen inputs where partial oxidation reactions provide
necessary heat without consuming hydrogen. As oxygen increases, each polymer
reaches a critical threshold where hydrogen production begins declining due to the
transition from gasification to combustion conditions.
The most striking differences appear in how each plastic responds to increasing oxygen.
PET shows the earliest decline beginning around 400 kg/hr air and reaches almost zero
production by ~900 kg/hr air, primarily due to its inherent oxygen content (~33% by
weight) that effectively serves as an internal oxidant. PS follows with decline starting at
~700 kg/hr O₂, limited by its aromatic structure with lower hydrogen availability.
HDPE and PP demonstrate superior resistance to oxidation, maintaining hydrogen
production up to ~1800-2000 kg/hr O₂ respectively, owing to their higher hydrogen
content and absence of oxygen in their molecular structure.

Effect on outlet CO2 flow on Inlet air flow


1800
1600
1400

CO2 flow 1200


1000
ps
800 pp
600 hdpe
pet
400
200
0
0 500 1000 1500 2000 2500
Inlet air flow

Figure 10 Flowrate of CO2 in the outlet by varying the inlet air flow rate

The graph illustrates the CO₂ flow rate (kg/hr) as a function of varying air flow rate for
different plastic feedstocks (PS, PP, HDPE, and PET). It demonstrates three distinct
phases: an initial low CO₂ production region at lower air flow rates, a transition phase
with rising CO₂ production, and a plateau region where maximum CO₂ production is
achieved. PET shows the earliest rise in CO₂ flow (~200-300 kg/hr) due to its inherent
oxygen content (~33% by weight), which acts as an internal oxidant during gasification.
This accelerates the shift from partial oxidation (C + ½O₂ → CO) to complete oxidation
(C + O₂ → CO₂), resulting in earlier CO₂ formation. PET also stabilizes at the lowest
plateau (~1150 kg/hr) because of its lower carbon content compared to other plastics.
PS achieves the highest plateau (~1650 kg/hr) because of its aromatic structure, which
contains the highest carbon density among the tested plastics. The aromatic rings in PS
contribute more carbon atoms per unit mass, leading to greater potential for CO₂
formation under complete oxidation conditions. PP and HDPE exhibit similar profiles,
with plateaus at ~1500 kg/hr, reflecting their similar chemical structures and carbon
contents. Their delayed rise in CO₂ production compared to PET indicates that they
require higher air flow rates to transition fully to combustion conditions.

Effect on outlet CO flow on Inlet air flow


1000
900
800
700
600
CO flow

500 ps
pp
400
hdpe
300 pet
200
100
0
0 500 1000 1500 2000 2500
Inlet air flow

Figure 11 Flowrate of CO in the outlet by varying the inlet air flow rate

The graph illustrates carbon monoxide (CO) flow rates from the gasification of different
plastic feedstocks (PS, PP, HDPE, and PET) as a function of increasing air flow. At low
air flow rates, all plastics show increasing CO production as oxygen availability enables
partial oxidation reactions (C + ½O₂ → CO), which are favored under oxygen-limited
conditions. PS exhibits the highest peak CO production (~950 kg/hr) due to its carbon-
rich aromatic structure, which provides more sites for partial oxidation reactions. PP and
HDPE display similar profiles with maximum CO production around 850-900 kg/hr,
reflecting their comparable polyolefin structures. PET demonstrates the most distinctive
behavior, with an earlier peak (around 300-400 kg/hr) and much faster decline
compared to other plastics.
As air flow increases beyond the optimal partial oxidation range, CO production
declines for all feedstocks because the process transitions from partial oxidation to
complete oxidation. The additional oxygen promotes the conversion of CO to CO₂ (CO
+ ½O₂ → CO₂) and direct oxidation of carbon to CO₂ (C + O₂ → CO₂). PET reaches
zero CO production earliest (around 900 kg/hr) due to its inherent oxygen content
(~33% by weight), which accelerates the transition to complete oxidation conditions.
The polyolefins (PS, PP, HDPE) continue producing CO until much higher air flow
rates (1500-2000 kg/hr), demonstrating their greater resistance to complete oxidation.

Effect on outlet CH4 flow on Inlet air flow


40

35

30

25
CH4 flow

20 ps
pp
15 hdpe
pet
10

0
0 500 1000 1500 2000 2500
Inlet air flow

Figure 12 Flowrate of CH4 in the outlet by varying the inlet air flow rate

The graph illustrates methane (CH₄) flow rates as a function of increasing air flow
during gasification of four different plastic feedstocks (PS, PP, HDPE, and PET).
Initially, PP demonstrates the highest methane yield (~35 kg/hr), followed by HDPE
(~30 kg/hr), PS (~18 kg/hr), and PET (~5 kg/hr), directly reflecting their hydrogen-to-
carbon ratios and molecular structures. PP and HDPE show superior methane generation
due to their high hydrogen content and lack of oxygen in their polyolefin chains, while
PS produces less methane due to its stable aromatic rings that resist methanation. PET
exhibits the lowest methane production and earliest decline (reaching zero around 500
air flow units) because its inherent oxygen content creates less favorable conditions for
methane formation and accelerates oxidation reactions.
As air flow increases, all plastics show characteristic declines in methane production,
with PP displaying a notable step-like decrease around 700-800 kg/hr of air. This
decline occurs because increasing oxygen promotes competing reactions including
direct methane oxidation (CH₄ + 2O₂ → CO₂ + 2H₂O) and steam reforming (CH₄ +
H₂O → CO + 3H₂), both of which consume methane. By approximately 1200 kg/hr of
air, methane production ceases entirely for all feedstocks as the system transitions
completely from reducing conditions (favorable for methane) to oxidizing conditions.

6.2. Effect of Gasifier temperature on the outlet composition

Effect of outlet H2 flow on Gasifier Temperature


80

70

60

50
H2 flow

40 ps
pp
30 hdpe
20 pet

10

0
400 600 800 1000 1200 1400 1600 1800 2000 2200
Temperature of Gasifier

Figure 13 Flowrate of H2 in the outlet by varying Gasifier Temperature

The graph depicts hydrogen flow rate (kg/hr) as a function of gasifier temperature for
different plastic feedstocks (PP, HDPE, PS, and PET), displaying a scientifically valid
relationship between thermal conditions and hydrogen production. Initially, at
temperatures below approximately 500°C, hydrogen production is minimal for all
plastic types because there's insufficient thermal energy to effectively break carbon-
hydrogen bonds in the polymer chains.
As the temperature increases from 500°C to approximately 1000-1200°C, all feedstocks
exhibit a rapid rise in hydrogen production due to enhanced thermal decomposition and
the increasing dominance of endothermic hydrogen-producing reactions like water-gas
(C + H₂O → CO + H₂) and steam reforming (CH₄ + H₂O → CO + 3H₂). The graph
correctly shows that PP achieves the highest maximum hydrogen yield (~72 kg/hr),
followed by HDPE (~67 kg/hr), PS (~42 kg/hr), and PET (~22 kg/hr), which directly
corresponds to their hydrogen content and molecular structures. PP and HDPE contain
more hydrogen per unit mass due to their simple hydrocarbon chains, while PS has less
hydrogen due to its aromatic rings, and PET has the lowest yield due to its oxygen-
containing ester groups that limit hydrogen availability.
The plateau behaviour observed above 1200-1500°C accurately represents the
thermodynamic limitations of the gasification process, where maximum conversion of
the available hydrogen in each plastic type has been achieved. This plateau formation is
consistent with literature findings that optimal hydrogen production typically occurs
within the 800-1000°C range, beyond which incremental gains become minimal or even
slightly negative as competing reactions begin to influence the equilibrium.
Effect of outlet CO flow on Gasifier Temperature

700

600

500

400
CO flow

ps
300 pp
hdpe
200 pet

100

0
400 600 800 1000 1200 1400 1600 1800 2000 2200
Temperaure of Gasifier

Figure 14 Flowrate of CO in the outlet by varying Gasifier Temperature

The graph illustrates carbon monoxide (CO) flow rate (kg/hr) as a function of gasifier
temperature for different plastic feedstocks (PS, PP, HDPE, and PET). Below
approximately 500°C, CO production is negligible for all plastics because there's
insufficient thermal energy to effectively break molecular bonds and drive gasification
reactions.
As temperature increases from 500°C to approximately 1000-1100°C, all feedstocks
exhibit a rapid rise in CO production due to accelerated thermal decomposition and the
promotion of critical endothermic reactions that favor CO formation, such as the
Boudouard reaction (C + CO₂ → 2CO) and partial oxidation (C + ½O₂ → CO). PET
demonstrates substantially higher CO production (plateauing at ~640 kg/hr) compared
to the other plastics (PS, PP, HDPE, which plateau around 350-380 kg/hr) because PET
contains significant oxygen in its molecular structure (-(C₁₀H₈O₄)-n), providing an
internal source of oxygen that facilitates CO formation even without external oxidants,
whereas PS, PP, and HDPE are pure hydrocarbons that require external oxygen sources
to form CO. The plateau behavior above approximately 1100°C correctly represents the
thermodynamic equilibrium state where maximum carbon conversion to CO has been
achieved, with further temperature increases providing minimal additional benefit to CO
formation.

Effect of outlet CO2 flow on Gasifier Temperature


400

350

300

250
CO2 flow

200 ps
pp
150
hdpe
100 pet

50

0
400 600 800 1000 1200 1400 1600 1800 2000 2200
Temperature of Gasifier

Figure 15 Flowrate of CO2 in the outlet by varying Gasifier Temperature

At lower temperatures (500-600°C), CO₂ production is significant, particularly for PET


which shows remarkably higher initial values (~350 kg/hr) compared to the other
plastics (~100-120 kg/hr) due to its inherent oxygen content (~33% by weight) that
facilitates CO₂ formation during initial decomposition. As temperature increases from
600°C to approximately 1100°C, all feedstocks exhibit a sharp decline in CO₂
production because higher temperatures increasingly favor the endothermic Boudouard
reaction (C + CO₂ → 2CO), which consumes CO₂ to produce CO.
This explains the corresponding rise in CO production observed in the previous graph,
as these reactions are thermodynamically linked. PET demonstrates the steepest decline
due to its oxygen-rich structure that promotes early decomposition, while the pure
hydrocarbons (PS, PP, HDPE) show more gradual decreases. By approximately 1000-
1100°C, CO₂ production approaches zero for all plastics as the gasification process
shifts almost entirely toward CO formation at these elevated temperatures. The overall
nature of this graph correctly represents the fundamental thermodynamic principles of
high-temperature gasification, where increased temperature favors products of
endothermic reactions (CO) over exothermic reaction products (CO₂).

Effect of outlet CH4 flow on Gasifier Temperature


200
180
160
140
120
CH4 flow

100 ps
80 pp
hdpe
60
pet
40
20
0
400 600 800 1000 1200 1400 1600 1800 2000 2200
Gasifier Temperature

Figure 16 Flowrate of CH4 in the outlet by varying Gasifier Temperature

The graph shows methane (CH₄) flow rate (kg/hr) as a function of gasifier temperature
for different plastic feedstocks (PS, PP, HDPE, and PET). At lower temperatures (below
500°C), CH₄ production is high for all feedstocks, with PP (~180 kg/hr) and HDPE
(~160 kg/hr) producing the most methane due to their high hydrogen content and simple
hydrocarbon structures. PS produces less methane (~80 kg/hr) because its aromatic
rings resist decomposition into CH₄, while PET generates the least (~20 kg/hr) due to
its oxygen-rich structure favoring CO and CO₂ formation instead of hydrocarbons.
As the temperature increases beyond 500°C, CH₄ production declines sharply for all
plastics because higher temperatures favor endothermic reactions like steam reforming
(CH₄ + H₂O → CO + 3H₂) and thermal decomposition (CH₄ → C + 2H₂), which
consume methane to produce syngas components like CO and H₂. By approximately
1000-1200°C, CH₄ flow approaches zero for all feedstocks as the equilibrium shifts
entirely toward syngas production. The steep decline is more pronounced for PP and
HDPE due to their higher initial methane yields, while PET exhibits a flatter curve
because of its inherently low methane production. This behavior aligns with established
gasification principles, where elevated temperatures progressively suppress hydrocarbon
formation in favor of gaseous products like CO and H₂.

6.3. Effect of Gasifier Pressure on Outlet composition

Effect of outlet H2 flow and outlet CO flow on Gasifier Pressure


80

70

60

50
H2 flow

40 ps
pp
30 hdpe
pet
20

10

0
0 5 10 15 20 25
Pressure of Gasifier

Figure 17 Flowrate of H2 in the outlet by varying Gasifier Pressure

700

600

500

400 ps
CO flow

pp
300 hdpe
pet
200

100

0
0 5 10 15 20 25
Pressure of Gasifier

Figure 18 Flowrate of CO in the outlet by varying Gasifier Pressure

The graphs illustrate carbon monoxide (CO) and hydrogen (H₂) flow rates (kg/hr) as
functions of gasifier pressure (1-20 bar) for different plastic feedstocks, and both
demonstrate decreasing trends. The CO flow graph shows PET producing substantially
higher CO (~610 kg/hr at 1 bar declining to ~380 kg/hr at 20 bar) compared to PS, PP,
and HDPE (~350-330 kg/hr declining to ~230-210 kg/hr) due to PET's significant
oxygen content (~33% by weight) that inherently promotes CO formation.
Similarly, the H₂ flow graph demonstrates PP generating the highest hydrogen yield
(~72 kg/hr at 1 bar declining to ~48 kg/hr at 20 bar), followed by HDPE (~65 kg/hr to
~45 kg/hr), PS (~40 kg/hr to ~28 kg/hr), and PET the lowest (~20 kg/hr to ~13 kg/hr),
directly correlating with their hydrogen content. These declining trends with increasing
pressure are thermodynamically correct and align with Le Chatelier's principle, as key
gasification reactions (C + H₂O → CO + H₂, CH₄ + H₂O → CO + 3H₂) increase the
number of gas molecules, making them less favorable at higher pressures. The
consistent negative slopes indicate the system is behaving according to established
gasification principles, where lower pressures generally favor syngas (CO + H₂)
production while higher pressures tend to favor products with fewer gas molecules, such
as methane and heavier hydrocarbons. The relative positions of each plastic type remain
consistent across the pressure range, confirming that feedstock composition remains the
dominant factor affecting product distribution even as pressure changes affect the
overall equilibrium.

Effect of CO2 flow and CH4 flow on Gasifier Pressure

180
160
140
120
CO2 flow

100
ps
80 pp
60 hdpe
pet
40
20
0
0 5 10 15 20 25
Pressure of Gasifier

Figure 19 Flowrate of CO2 in the outlet by varying Gasifier Pressure


80
70
60
50

CH4 flow
40 ps
30 pp
hdpe
20 pet
10
0
0 5 10 15 20 25
Pressure of Gasifier

Figure 20 Flowrate of CH4 in the outlet by varying Gasifier Pressure

The graphs depict methane (CH₄) and carbon dioxide (CO₂) flow rates as functions of
gasifier pressure (1-20 bar) for different plastic feedstocks, and both demonstrate
scientifically increasing trends. The CH₄ flow graph shows PP producing the highest
methane yield (~70 kg/hr at 20 bar), followed by HDPE (~60 kg/hr), PS (~32 kg/hr),
and PET the lowest (~9 kg/hr), which directly correlates with their hydrogen content
and molecular structure.
This increasing trend with pressure follows Le Chatelier's principle, as the methanation
reaction (C + 2H₂ → CH₄) reduces the number of gas molecules, making it
thermodynamically favored at higher pressures. Similarly, the CO₂ flow graph shows
PET generating dramatically higher CO₂ (~150 kg/hr at 20 bar) compared to PS, PP,
and HDPE (~40-50 kg/hr), due to PET's significant oxygen content (~33% by weight)
that inherently promotes CO₂ formation. The water-gas shift reaction (CO + H₂O ⇌
CO₂ + H₂) becomes increasingly favorable at higher pressures in this complex system,
particularly for oxygen-containing feedstocks. These pressure-dependent behaviours
directly oppose the trends seen in H₂ and CO production (which decrease with pressure)
because gasification naturally shifts toward products with fewer gas molecules (CH₄,
CO₂) rather than more gas molecules (H₂, CO) as pressure increases.
6.4. Effect of Steam on Gasification Products

80
70
60
50
H2 flow rate

40 0 kg/hr
100 kg/hr
30 200 kh/hr
20 300 kg/hr

10
0
400 600 800 1000 1200 1400 1600 1800 2000 2200
Temperature of Gasifier for Polystyrene Gasification

Figure 21 Effect of H2 flow on Introduction of steam

The graph illustrates the profound impact of steam addition on hydrogen production
during plastic gasification, showing that increasing steam flow rates (from 0 to 300
kg/hr) significantly enhances hydrogen yield across all gasification temperatures. As
temperature rises from 500°C to approximately 1000°C, hydrogen production increases
dramatically for all steam flow rates, with the 300 kg/hr steam flow achieving the
highest hydrogen yield (~75 kg/hr) compared to no steam addition (~42 kg/hr) at
plateau. This enhancement occurs because steam facilitates critical hydrogen-producing
reactions including steam reforming (CH₄ + H₂O → CO + 3H₂), water-gas shift (CO +
H₂O → CO₂ + H₂), and water-gas reactions (C + H₂O → CO + H₂). However, steam
addition also introduces several negative effects to the gasification process: it requires
significant energy input for steam generation, potentially reducing overall process
efficiency.
I has been found that the introduction of steam for the process increases the heat duty of
the gasifier by two times. In the simulation of 625 kg/hr of waste plastics, for every 100
kg/hr of increase in the water content to the gasifier, the heat duty increased two folds.
For 100 kg/hr of water the heat duty was 593.7 kcal/s, for 200 kg/hr of water flow, the
heat duty was 1061.9 kcal/s and for 300 kg/he of water flow, the heat duty was 1532.6
kcal/s.

Effects of Nitrogen on Gasification Process


The presence of nitrogen in gasification systems significantly impacts process efficiency
and product distribution. Nitrogen acts as an inert diluent that reduces the partial
pressure of reactive components in the gasifier, thereby diminishing the thermodynamic
driving force for key reactions. This dilution effect decreases the concentration of
hydrogen and carbon monoxide in the product gas, reducing both yield and calorific
value. Additionally, the presence of nitrogen increases the volumetric flow through
process equipment, necessitating larger vessel dimensions and downstream separation
units, which substantially raises both capital and operational expenditures.

Routing the Moisture stream to Gasifier

The recirculation of moisture to the gasifier, specifically at 20% of the weight of input
plastics, represents an optimal approach for maximizing hydrogen production and
overall process efficiency. This moisture content aligns precisely with the recommended
range (10-20%) for achieving high calorific value in produced gas, as identified in
literature[9]. The graph clearly demonstrates that increasing steam flow significantly
enhances hydrogen production, with hydrogen yield rising from approximately 42 kg/hr
without steam to 75 kg/hr at the highest steam flow rate (300 kg/hr). This substantial
78.6% increase in hydrogen production occurs because moisture participates in key
hydrogen-producing reactions including steam reforming (CH₄ + H₂O → CO + 3H₂),
water-gas shift (CO + H₂O → CO₂ + H₂), and water-gas reactions (C + H₂O → CO +
H₂).
The 20% moisture level strikes an optimal balance, sufficient to drive hydrogen-
producing reactions without the thermal penalties associated with excessive moisture
that could reduce gasification temperature and performance.

Figure 22 Modified Flowsheet for an optimised approach of production of hydrogen by routing the
moisture to the gasifier
7. CONCLUSION

This study investigated the thermochemical conversion of waste plastics to hydrogen


through gasification using Aspen Plus simulation. The process began with a feed rate of
625 kg/hr of various plastic types (HDPE, PP, PS, and PET), which was dried to 500
kg/hr before undergoing decomposition and pyrolysis. The comprehensive parametric
analysis revealed significant insights into the effects of plastic composition, process
conditions, and reaction pathways on product yields and distributions.

The comparative evaluation of different plastic feedstocks demonstrated their inherent


suitability for hydrogen production based on their molecular structures. Polypropylene
(PP) consistently yielded the highest hydrogen production (approximately 67-72 kg/hr
at optimal conditions), followed closely by high-density polyethylene (HDPE) at 60-65
kg/hr. This superior performance can be attributed to their favorable hydrogen-to-carbon
ratios and simple hydrocarbon structures. Polystyrene (PS) produced significantly less
hydrogen (approximately 38-42 kg/hr) due to its aromatic rings that reduce hydrogen
availability and increase carbon deposition tendencies. Polyethylene terephthalate (PET)
exhibited the lowest hydrogen yields (approximately 19-22 kg/hr) owing to its oxygen-
containing structure that promotes CO and CO₂ formation at the expense of hydrogen.
These results confirm that polymer chemistry fundamentally determines the theoretical
hydrogen production potential, with polyolefins (PP and HDPE) representing the most
promising feedstocks for hydrogen-focused gasification.

Temperature emerged as a critical parameter governing product distribution and


hydrogen yield. At low temperatures (below 600°C), hydrogen production remained
minimal across all feedstocks due to insufficient thermal energy to break carbon-
hydrogen bonds effectively. As temperatures increased from 600°C to approximately
1000-1200°C, hydrogen production rose dramatically, reaching maximum values of 75
kg/hr with steam addition at 900-1000°C. This temperature-dependent behavior reflects
the dominance of endothermic reactions like steam reforming (CH₄ + H₂O → CO +
3H₂) and water-gas (C + H₂O → CO + H₂) at elevated temperatures. The observed
plateauing effect above 1200°C indicates that thermodynamic equilibrium limits are
reached, beyond which additional thermal energy provides diminishing returns.
Simultaneously, methane production exhibited an inverse relationship with temperature,
decreasing sharply from initial values of 160-180 kg/hr at 500°C to near-zero at 1200°C
as reforming reactions progressively converted methane to hydrogen and carbon
monoxide.

The gasifier pressure demonstrated significant influence on product distribution, with


clear trends that aligned with Le Chatelier's principle. As pressure increased from 1 to
20 bar, hydrogen production decreased substantially for all plastics—from 72 to 48
kg/hr for PP, 65 to 45 kg/hr for HDPE, 40 to 28 kg/hr for PS, and 20 to 13 kg/hr for
PET. Carbon monoxide followed a similar declining pattern with increased pressure.
Conversely, methane and carbon dioxide production increased with pressure, reflecting
the thermodynamic preference for reactions that reduce the total number of gas
molecules at higher pressures. This behaviour confirms that lower operating pressures
(1-5 bar) are optimal for maximizing hydrogen yield, while higher pressures favour
methanation reactions that consume hydrogen.

The effect of air flow rate (oxygen availability) revealed complex but scientifically
consistent behavior across all plastic types. At low air flow rates, hydrogen production
remained stable as partial oxidation reactions provided necessary heat while minimizing
hydrogen consumption. As air flow increased beyond critical thresholds specific to each
plastic type, hydrogen production declined precipitously due to the transition from
gasification to combustion conditions. PET showed the earliest decline, beginning
around 400 kg/hr of air and reaching zero production by approximately 900 kg/hr, while
PP demonstrated the greatest resistance to oxidation, maintaining hydrogen production
up to approximately 2000 kg/hr of air. This behavior underscores the critical importance
of oxygen control in gasification processes, where excess oxygen rapidly shifts the
process toward complete combustion products rather than hydrogen.
8. FUTURE SCOPE

 Include various reactor such as plug flow reactors (PFRs), Continuous Stirred Reactors
(CSTRs) for modelling the process.
 Study the gasification of mixed plastic waste or co-gasification with biomass to simulate
real-world waste streams.
 Explore heat recovery and integration options to improve the overall energy efficiency of
the process.
 Perform a techno-economic assessment to determine the commercial viability and
identify cost reduction opportunities.
9. REFERENCES

1 Han, J., Choi, Y., & Kim, J. (2020). Development of the process model and optimal drying
conditions of biomass power plants. ACS Omega, 5(6), 2811–2818.
https://doi.org/10.1021/acsomega.9b03557
2 Kannan, P., Shoaibi, A. A., & Srinivasakannan, C. (2012). Process simulation and
sensitivity analysis of waste plastics gasification in a fluidized bed reactor. WIT
Transactions on Ecology and the Environment. https://doi.org/10.2495/wm120171
3 Encinar, J., & González, J. (2008). Pyrolysis of synthetic polymers and plastic wastes.
Kinetic study. Fuel Processing Technology, 89(7), 678–686.
https://doi.org/10.1016/j.fuproc.2007.12.011
4 Yao, D., Yang, H., Chen, H., & Williams, P. T. (2018). Co-precipitation, impregnation and
so-gel preparation of Ni catalysts for pyrolysis-catalytic steam reforming of waste plastics.
Applied Catalysis B Environment and Energy, 239, 565–577.
https://doi.org/10.1016/j.apcatb.2018.07.075
5 Kannan, P., Al, A., & Srinivasak, C. (2012). Optimization of waste plastics gasification
process using Aspen-Plus. In InTech eBooks. https://doi.org/10.5772/48754
6 Ahmad, M. (2009, January 1). Kinetic modeling of biomass steam gasification system for
hydrogen production with CO2 adsorption.
https://www.academia.edu/1759616/Kinetic_modeling_of_biomass_steam_gasification_s
ystem_for_hydrogen_production_with_CO2_adsorption
7 Sravani, P., Povari, S., Alam, S., Nakka, L., Srinath, S., & Chenna, S. (2024). Co-
gasification of Waste Biomass and Plastic for Syngas Production with CO2 Capture and
Utilization: Thermodynamic Investigation. Longdom. https://doi.org/10.35248/2157-
7544.24.15.379
8 Bashir, M. A., Ji, T., Weidman, J., Soong, Y., Gray, M., Shi, F., & Wang, P. (2025). Plastic
Waste Gasification for Low-Carbon Hydrogen Production: A Comprehensive Review.
Energy Advances, 4(3), 330–363. https://doi.org/10.1039/d4ya00292j
9 Shah, H. H., Amin, M., Iqbal, A., Nadeem, I., Kalin, M., Soomar, A. M., & Galal, A. M.
(2023). A review on gasification and pyrolysis of waste plastics. Frontiers in Chemistry,
10. https://doi.org/10.3389/fchem.2022.960894
10 Dhrioua, M., Ghachem, K., Hassen, W., Ghazy, A., Kolsi, L., & Borjini, M. N. (2022).
Simulation of biomass air gasification in a bubbling fluidized bed using Aspen Plus: a
comprehensive model including TAR production. ACS Omega, 7(37), 33518–33529.
https://doi.org/10.1021/acsomega.2c04492
11 Gao, Y., Wang, M., Raheem, A., Wang, F., Wei, J., Xu, D., Song, X., Bao, W., Huang, A.,
Zhang, S., & Zhang, H. (2023). Syngas Production from Biomass Gasification: Influences
of Feedstock Properties, Reactor Type, and Reaction Parameters. ACS Omega, 8(35),
31620–31631. https://doi.org/10.1021/acsomega.3c03050

You might also like