You are on page 1of 15

Process Safety and Environmental Protection 184 (2024) 907–921

Contents lists available at ScienceDirect

Process Safety and Environmental Protection


journal homepage: www.journals.elsevier.com/process-safety-and-environmental-protection

Synergistic transformation: Kinetic and thermodynamic evaluation of


co-pyrolysis for low-rank bituminous coal and polyurethane foam waste
Zaid Mohammad Banyhani a, Waqar Ul Habib Khan a, Hala H. Abd El-Gawad b,
Mustafa Anwar a, *, Asif Hussain Khoja a, *, Muhammad Hassan a, Rabia Liaquat a,
Zeinhom M. El-Bahy c
a
U.S-Pakistan Centre for Advanced Studies in Energy (USPCAS-E), National University of Sciences & Technology (NUST), Sector H-12, Islamabad 44000, Pakistan
b
Department of Chemistry, Faculty of Science and Arts, King Khalid University, Mohail, Assir, Saudi Arabia
c
Department of Chemistry, Faculty of Science, Al-Azhar University, Nasr City, 11884 Cairo, Egypt

A R T I C L E I N F O A B S T R A C T

Keywords: Co-pyrolysis is regarded as a sustainable and ecologically friendly method of improving waste management,
Co-pyrolysis pollution control, and renewable energy security. This investigation reports a thermo-kinetic study of low-rank
Thermo-Kinetics coal-polyurethane foam waste (C-PU) blends through co-pyrolysis and was characterized using a number of
Coats-Redfern method
techniques such as CHNS-O, GCV, and FTIR. The thermal degradation behavior of C-PU blends during co-
Low-Rank Bituminous Coal
Plastic waste
pyrolysis via TGA is studied, whereas the degradation process happens in three stages i.e. moisture, devolatili­
Polyurethane foam zation, and slow degradation of the char in the temperature ranges viz. (25 –140 ◦ C), (140 – 580 ◦ C), and (580 –
Reactivity analysis 900 ◦ C) respectively. Synergistic effects observed may improve product qualities compared to those from
separate pyrolysis. The presence of synergistic effects during co-pyrolysis was demonstrated by the positive
variation in WL% and DTGmax values. The Coats-Redfern integral method was utilized to investigate the thermo-
kinetic parameters of C-PU blends through twelve (12) reaction mechanism models. The activation energy (Ea)
for 100C and 100PU was 60.08 kJ/mol and 100.8 kJ/mol via F2 and F3 models, although the optimum blend
(50 C-50PU) showed 67.01 kJ/mol and 26.74 kJ/mol via F2 and D3 model in the first and second stage.
Thermodynamic characteristics i.e. (ΔG) and (ΔH) displayed positive values, while (ΔS) for each C–PU blend
was negative, but the best blend further lowered it. It was found that the fuels could be ranked in the following
order 100PU> 50 C-50PU> 100 C based on mean reactivity and pyrolysis factor. The co-pyrolysis of poly­
urethane foam waste with low-rank coal to develop alternative energy sources has been proposed as a potential
method and is critical for designing an effective large-scale reactor system thus understanding the solid-state co-
pyrolytic kinetics.

declining credibility (Cai et al., 2008). However, only 9% of its produced


1. Introduction waste worldwide was reprocessed, and about 80% accumulated and
ended up in landfills (Evode et al., 2021), thus it is vital to investigate
Plastics in the last 50 years played a dynamic role in improving alternate plastic recycling solutions. To decrease the dissipation of fossil
human livelihood owing to their low cost, and the world has reached fuels and their accompanying releases, there is an imperative necessity
around 300 million tonnes of products, that’s why the accretion of non- for renewable energy incorporation and the rapid growth in the econ­
degradable plastic wastes has become a major environmental burden omy necessitates a large amount of energy, the majority of which is still
(Wang et al., 2019). They are broadly engaged in a range of various derived from fossil fuels. However, the usage of fossil fuels causes
fields such as industry, agriculture, electricity and electronics, and the several issues, such as pollution, non-renewability, and so on (Zhou
automobile industry (Pinto et al., 1999), posing a danger to the envi­ et al., 2019; Singh et al., 2022).
ronment and natural resources in the form of greenhouse gases and Worldwide, polyurethane foam (PU) is among the most used poly­
harmful pollutants released by incineration (Zhang et al., 2021a; Wu mers that are utilized in rigid or flexible forms for insulation, automotive
et al., 2022), which is trailing significance due to operational costs and purposes, and structural materials. The isocyanates and the polyols are

* Corresponding authors.
E-mail addresses: mustafa@uspcase.nust.edu.pk (M. Anwar), asif@uspcase.nust.edu.pk (A.H. Khoja).

https://doi.org/10.1016/j.psep.2024.01.041
Received 1 September 2023; Received in revised form 6 January 2024; Accepted 9 January 2024
Available online 20 January 2024
0957-5820/© 2024 Institution of Chemical Engineers. Published by Elsevier Ltd. All rights reserved.
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Nomenclature Pf Pyrolysis Factor.


R2 Linear Regression.
A Pre-exponential Factor. R Gas Constant.
100 C/C Low-rank coal. Rm Relative Mean Reactivity.
CHNS-O Carbon, Hydrogen, Nitrogen, Sulfur, and Oxygen/Ultimate TG/TGA Thermo-gravimetric Analysis.
Analysis. RL% Residue Left %.
C-PU Low-rank coal-Polyurethane foam. WL% Weight Loss %.
DTGmax Maximum Derivative Thermogravimetry. α Conversion Factor.
DTG Derivative Thermogravimetry. β Heating Rate.
Ea Activation energy. ΔH Change in Enthalpy.
FTIR Fourier Transform Infrared Spectroscopy. ΔG Change in Gibbs free energy.
GCV Gross Calorific Value. ΔS Change in Entropy.
H/C Hydrogen-to-carbon atomic ratio. f(α) Reaction Mechanism.
h Planck Constant. g(α) Integral Function.
K Rate Constant. kb Boltzmann Constant.
O/C Oxygen-to-Carbon atomic ratio. (wt%/◦ C) Weight Loss % per degree centigrade/DTG unit.
100PU/PU Polyurethane foam. Tp Peak Degradation Temperature.

the primary building blocks of PU and more than 150 types of foam can technology may include waste reduction, chemical recovery, and an
be produced by selecting different combinations (Crescentini et al., alternative in widely recognized industrial transformation methods and
2019; Simón et al., 2018), and used as an important copolymer that is contributing to environmental protection (Melendi-Espina et al., 2015).
widely used in manufacturing, the yearly production since 2015 has Coal and plastic wastes exhibit an abundance of diverse applications,
reached 16 million tons per year, majority of this is not biodegradable, and low costs in numerous industries (Hassan et al., 2016).
so only a small amount is recycled. However, the transformation of PU Various studies on the co-blending of coal and plastics have been
into useful fuels such as gas, liquids, and char by thermochemical con­ conducted (Saha et al., 2018), and better results were observed due to
version i.e. pyrolysis an utmost ecologically sustainable approach to synergy at a high-temperature region to give better results than the
take care of its disposal issue for further utilization (Yao et al., 2020a), pyrolysis of coal (Sharma and Ghoshal, 2010). Zhang et. al (Zhang et al.,
and upgrade the plastic waste since the organic compounds in the plastic 2022a) studied the co-pyrolysis of Chinese low-rank coal with
be wasted (Jin et al., 2019). Several researchers studied the sole pyrol­ high-density polyethylene (HDPE) which revealed that a better quality
ysis process of polyurethane foam (Jomaa et al., 2015; Stančin et al., of tar is obtainable by charging the coal on top of plastic during the
2019), but it yields a lot of ash, and the oil has a lot of oxygen, viscosity, co-pyrolysis process. Hong et. al studied the co-pyrolysis behavior of
corrosiveness, and moisture in it (Seah et al., 2023). Previous re­ polyvinyl chloride (PVC) with coal via ReaxFF simulations and their
searchers have demonstrated two pyrolysis sub-reactions for the poly­ results revealed that PVC encourages CH2O, CO, and CO2 generation
urethane foam degradation in a nitrogen environment described as while it inhibits H2O generation (Hong et al., 2022). Co-pyrolysis in­
follows (Li et al., 2016): teractions of low-rank coal and polystyrene were studied for the in-situ
First Reaction (foam degradation) in Eq. (1): detection of primary volatiles and their kinetic study revealed that
effective supplements to the co-pyrolysis interactions were provided by
Polyurethane foam→Melt (Polyol) + Gas(Isocyanate) (1)
the change in activation energy (Wu et al., 2021). Meanwhile, the
Second Reaction (melt degradation) in Eq. (2): co-pyrolysis of low-rank coal with polyethylene(PE) revealed that PE
encourages the breaking of C-C bond and inhibits C-O bond breaking to
Melt(Polyol)→Residue + Gas(Polyol, H2 CO, H2 O, and CH 4 ) (2)
enhance H2 production and inhibit the generation of oxygen-related
Coal is made up of an organic macromolecular skeletal matrix gasses (Feng et al., 2023). Several authors have explored the impact of
structure and low-molecular-weight chemicals and its combustion gen­ synergistic effects among low-rank coal and various materials such as
erates thermal energy resulting in hydrogen molecules and continuous sawdust and hemp (Gohar et al., 2022), rice husk (Tauseef et al., 2022),
free radicals (Jayanti et al., 2012; Rajasekhar Reddy and Vinu, 2018). It and algae consortium (Khan et al., 2022). Recently various authors have
maintains its high calorific value and is a significant source of heat and carried out thermo-kinetic studies of coal with other materials such as
electricity generation. Its usage encounters approximately 29% of the wheat straw (Zhou et al., 2019), Scenedesmus microalgae (Nyoni et al.,
basic world energy consumption and 39% of global power demand 2020), and sugarcane bagasse (Saeed et al., 2021a) using the
(Mallick et al., 2018). However, over time it has become less popular coats-redfern integral method which demonstrated excellent results.
except for nations with high coal reserves (Gouws et al., 2021). In many Similarly, several researchers have investigated the synergistic effect
nations, coal is a primary source of energy production, nevertheless, its and thermo-kinetic analysis of the co-pyrolysis of polyurethane and
reserves are rapidly depleting, also, its pyrolytic products are low and its biomass such as sawdust mixture (Stančin et al., 2021). As a result, the
carbon emissions are high owing to a lower H/C ratio (Nyoni et al., novelty of this work is that, for the first time, both coal and polyurethane
2020). Transforming co-blends to liquid hydrocarbon fuels has greater foam have been examined as a fuel blend to acquire suitable deductions
commercialization potential than other options decreasing the carbon about the feedstock appropriateness for the process, which was sup­
imprint in the environment but also allowing the use of locally available ported by an examination of favorable procedure settings.
low-rank coals (Folkedahl et al., 2011). In this present work, the co-pyrolysis behavior of C-PU blends was
To overcome these issues co-blending of coal and plastic waste has explored for the first time in a thermogravimetric analyzer (TGA) under
become an attractive pursuit as feedstocks for the co-pyrolysis process, an inert atmosphere. The aim was to find out how the thermogravi­
which is a promising technique due to its versatility in product gener­ metric, synergistic effects, and thermo-kinetic analysis of C-PU blends
ation (Wei et al., 2022; Khalid et al., 2024). The co-pyrolysis of waste were carried out to evaluate the capability of nominated feedstock for
plastic with fossil fuels is now being explored thus partially substituting co-pyrolysis. The reactivity analysis was carried out on parent fuels and
fossil fuels with higher-value fuel, particular advantages of this the optimum blend to compare and evaluate them. The impact of plastic

908
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Fig. 1. Schematic representation of C-PU blends preparation.

content on product quality and quantity is assessed by detecting the 2.2. Characterization of C-PU blends
synergistic effects. The thermo-kinetic factors for the total volatiles
released from C-PU blend fuels were deduced using the Coats–Redfern To analyze the percentage of carbon, hydrogen, and nitrogen in each
integral method via understanding twelve reaction mechanistic models sample of the prepared coal and polyurethane foam (C-PU) blends by
to have a better knowledge of the coal and polymer co-pyrolysis kinetic comparing them with the standard values, each sample weighed 80 mg
mechanism. The findings of this study can be utilized to assess the with a particle size of 0.2 mm which was folded in a prepared aluminium
compatibility of feedstock for co-pyrolysis and, more importantly, to foil cup and inserted into the disc in the CHN Analyzer (5ECHN2200,
deliver a thorough understanding of selecting optimal procedure vari­ CKIC, China). Sulfur Analyzer 5E-IRS II (CKIC, China) was used to
ables such as mixing ratio, heating rate, and the final temperature to determine the S content in C-PU blends. A weight of 300 mg was taken
produce alternative fuels. with a particle size of 0.2 mm for the Sulfur analyzer. To perform the
ultimate analysis, the percentage was determined by using the (N2, O2,
2. Materials and methods and He) gases (Khan et al., 2022). The customary procedure (ASTM
D5865–13) was followed and a Parr 6200 isoperibol oxygen bomb
2.1. Preparation of samples calorimeter was employed for analysis of the gross calorific value (GCV)
for each prepared sample of C-PU blends. Cary 630, Agilent Technolo­
The polyurethane foam (100PU) and low-rank coal (100 C) were gies (USA) was utilized for FTIR analysis to study the functional group in
collected locally and were used to prepare the blends. Fig. 1 demon­ C-PU blends and 0.4 g weight from each sample was used for the test. A
strates the graphic representation of C-PU blends. Primarily the pure 2 cm− 1 resolution with 4000 to 650 cm− 1 scanned range of infrared
materials were placed in an oven at 108 ◦ C for 24 h to confiscate the spectra was for the absorption characteristic.
moisture content. A 0.2 mm fine particle size was obtained by crushing
and sieving the raw materials through the Hard Grove Grindability
Index Tester (USA) and WS Tyler RX-29–10 (USA). The C-PU blends 2.3. Thermo-gravimetric analysis (TGA)
were prepared on a mass percent basis by mixing low-rank coal with
polyurethane foam with the ratios of (75 C-25PU) which represent The thermo-gravimetric analyzer (TGA) 5500 (TA Instruments, USA)
75 wt% of coal was mixed with 25 wt% of Polyurethane foam, corre­ was utilized to perform the thermal analysis of C-PU blends, each sample
spondingly, (50 C-50PU) and (25 C-75PU) blends were prepared weighed 10 mg to perform the thermo-gravimetric (TG/DTG) analysis.
accordingly. The Digital GSM Balance was utilized for weighing and Before experimentation, each one sample was purged for 30 min with N2
ensuring the blend preciseness of the C-PU blends. The Vortex mixer and having a flow rate of 10 mL min− 1, then the experiment was per­
(F20220176, VELP SCIENTIFICA, EUROPE) was then utilized for the formed, and then each sample was heated at a rate of 10 ◦ C/min from a
homogenization of the blends (Khan et al., 2022). temperature of 25 – 900 ◦ C and the WL% of C-PU blends were docu­
mented as a function of time and temperature under customary

909
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Table 1
Normally used reaction models’ algebraic expressions.
Reaction Model Reaction Mechanism Symbols 1 dα Integral form g(α) ¼ kt
Differential form f (α) =
k dt

Reaction Order Models First order F1 (1 − α) − ln(1 − α)


Second order F2 (1 − α)2 (1 − α)− 1 − 1
Third order F3 (1 − α)3 1[ ]
(1 − α)− 2 − 1
2
Fourth order F4 (1 − α)4 1[ ]
(1 − α)− 3 − 1
3
Power Law Power law; P2 P2 1 1
2(α)2 (α)2
Diffusivity Models Parabolic law D1 1 α2
α
2
Valansi equation D2 − [ln(1 − α) ]− 1 α + [(1 − α)ln(1 − α) ]
Jander equation D3 2[ 1] [ 1]
− 1 2
2(1 − α)3 1 − (1 − α)3 1 − (1 − α)3
Ginstling Brounstein equation D4 1 2
3[ ]− 1 2
(1 − α) 3 − 1 α − (1 − α)3

1−
2 3
Geometric Contraction Models Contracting disk R1 1 α
Contracting cylinder R2 1 1
2(1 − α)2 1 − (1 − α)2
Contracting sphere R3 2 1
3(1 − α)3 1 − (1 − α)3

conditions. The derivative WL% to temperature gives the differential 2.4. Thermo-Kinetic analysis
thermogravimetric (DTG) (wt%/◦ C). The thermo-kinetic parameters of
C-PU blends are calculated by analyzing the information acquired from The presence of various mechanisms makes studying co-pyrolytic
TG/DTG. kinetics a very complicated method (Gohar et al., 2022). Co-pyrolytic
kinetics helps optimize, operate, and design the process (Ali et al.,
2.3.1. Synergistic Effects Analysis 2018). The kinetic analysis for solid fuels namely low-rank coal (100 C)
The analysis of the interaction between 100PU and 100C during co- and polyurethane foam (100PU) is carried out by formulating the kinetic
pyrolysis is explained by utilizing the synergistic effect. Existing among equations to understand the mechanism of kinetics, and to get effective
the blends and determination of the effect by equating the theoretical kinetic factors namely activation energy (Ea), linear regression (R2), and
values of blends with the experimental values obtained from TG/DTG pre-exponential factor (A). The model-fitting method is ordinarily
was explored via the additive mode. The theoretical values obtained by employed, ascribing to its accessibility of assessing the kinetic factors
the additive mode supposed that no interactions happen amid 100C and directly. Various scientific researchers utilized the non-isothermal
100PU through co-pyrolysis (Park et al., 2010). The theoretical values coats-Redfern integral method to assess the kinetic behavior
can be estimated using Eq. (3): throughout co-pyrolysis. To acquire the total volatile release kinetics for
co-pyrolysis, this approach was applied to the Thermo-gravimetric
yC = XC (Yexp.C ) + XPU (Yexp.PU ) (3)
analysis in this study. Various reaction models have been proposed
(Gohar et al., 2022; Parthasarathy et al., 2021; Hameed et al., 2018) and
Where XC and XPU symbolize the weight ratio of 100C and 100PU in the
kinetic factors can be deduced based on these models. Almost twelve
blends correspondingly. Yexp.C and Yexp.PU represents the experimental
reaction models were utilized to establish the kinetic factors as
values of 100C and 100PU obtained from TG/DTG. For negative syn­
mentioned in Table 1. Since a larger percentage of WL% arises in the
ergistic effects, the experimental values will be lower than the theoret­
active pyrolysis area, the kinetic study was executed on this region
ical values indicating the inhibiting effect of co-pyrolysis, and for
(Naqvi et al., 2019), moisture stage was not included in the calculation.
positive synergistic effects, the experimental values will be greater than
In a single-step reaction, the thermal degradation of C-PU blends
the theoretical values, which designates the promoting impact of co-
forms volatiles (gases, tar) and char (solid residue) which can be
pyrolysis. The theoretical values are a combination of the values ob­
described in Eq. (5) (Yao et al., 2020b; Pan et al., 2021):
tained from individual samples according to their blending ratio of
weight (Khan et al., 2022). A→B + C (5)
Eq. (4) was utilized for the measurement of deviance among the
A general expression of the rate equation for the non-isothermal ki­
calculated and experimental values of (RL%), ((DTG)max %), and (WL
netics for the solid degradation is illustrated as follows in Eq. (6):
%). It will also highlight whether the synergistic effect exists or is absent
in the (C-PU) blends during co-pyrolysis (Gohar et al., 2022): da
= k(T) • f (a) (6)
YE − Yc dt
Deviation(%) = × 100 (4)
Yc Where the conversion rate is represented by (α) having values between
(0− 1), the reaction mechanism is f(α), and the rate constant is k (T). The
Where the calculated values obtained from Eq. (3) are represented by
sample’s mass is weighed continuously as a function of time and tem­
YC, and the experimental values acquired through TG/DTG are repre­
perature during the TG experimentation. Eq. (7), gives the quantity of
sented by YE. The larger the deviance amid the experimental and theo­
matter converted (α) at any time (t) throughout co-pyrolysis into vola­
retical values, the greater will be the synergistic effects’ value (Ding
tiles, calculated with mass range:
et al., 2021).
mi − mo
α= (7)
mi − mf

Where (mi ) is the initial mass of the material in (kg) before degradation,

910
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Table 2
Ultimate Analysis of Parent Fuels.
Sample Name Carbon (C%) Hydrogen (H%) Nitrogen (N%) Sulfur (S%) Oxygen (O%)a Oxygen to Carbon Ratio (O/C) Hydrogen to Carbon Ratio (H/C)

100 C 68.91 3.82 1.28 0.48 25.51 0.278 0.661


100PU 69.30 6.04 6.29 0.00 18.37 0.199 1.038
a
by difference

and the instantaneous mass of the material in (kg) is donated by (mo ) at 2019). The function g(α) is affected by the governing reaction mecha­
time t. through degradation and final mass (mf ) in (kg) after the nism, as well as the shape and size of the reacting particles (Liu et al.,
degradation of the sample (Zhang et al., 2021b). Additionally, Eq. (8) 2015).
gives the Arrhenius for measuring the rate constant k. Table 1 gives the kinetic factors (A and Ea) of the selected most
( ) suitable g(α) which were utilized for measuring the thermodynamic
Ea
k(T) = A • exp − (8) parameters i.e. change in Entropy (ΔS, kJ/mol K), change in Enthalpy
RT (ΔH, kJ/mol), and change in Gibbs free energy (ΔG, kJ/mol). These
properties can be determined from the kinetic factors values, using Eq.
Where the absolute temperature (K) is denoted by T, the activation
(16) and Eq. (17) given below (Tauseef et al., 2022):
energy (kJ/mol) is termed as Ea, the universal gas constant (8.314 J.K− 1.
mol− 1) is represented by R, and in the Arrhenius equation the pre- ΔH = Ea − R • TP (16)
exponential factor (min− 1) is denoted by A. The rate of a reaction is ( )
defined as the differential of conversion to time (Yao et al., 2022). Eq. Kb Tp
ΔG = Ea + R•Tp ln (17)
(9) is the reorganized form of the fundamental Arrhenius rate expression hA
at a constant β.
( ) Where the peak degradation temperature is denoted as Tp, having units
dα A
= • exp −
Ea
• f (α) (9) of kelvin (K), the Planck constant is represented by h, having a value of
dT β RT 6.63 × 10− 34 Js, the universal gas constant is expressed as R, and the
It is believed that for the majority of the time, exclusively associated Boltzmann constant is termed as Kb having a value of 1.38 × 10− 23 J/K.
with degradation reactions, the fuel co-pyrolysis is a reaction of first- Eq. (18) is utilized for measuring the entropy:
order (n = 1). Eq. (10) could be produced by reorganizing Eq. (9) at a ΔH − ΔG
constant β supposing that n = 1. ΔS = (18)
Tp
( )
dα A
= • exp −
Ea
• (1 − α) (10) The investigation of the Coats-Redfern integral technique gave the
dT β RT values of activation energy (Ea) via kinetic factors and these values of Ea
Eq. (11) is formed by incorporating the aforementioned equation: gave the values of ΔS, ΔG, and ΔH.

∫α
dα A
∫ T (
Ea
) 2.5. Reactivity Analysis
g(α) = = exp − dT (11)
f (α) β T0 RT
0 By using the TG and DTG curves data the Relative mean reactivity
g(α) symbolizes the reaction model’s integrated form in Eq. (6). Eq. (Rm) and pyrolysis factor (Pf) for 100 C, 100PU, and their best blend
(12) is formed after the utilization of the Taylor series by The Coats- 50 C-50PU was calculated. The sum of all of the peak heights to the
Redfern integral method in Eq. (10): corresponding temperature is an indication of the mean reactivity of the
[ ] [ ( )] pyrolytic biomass. Relative mean reactivity (Rm) is the ratio of the
ln
g(α)
= ln
AR
1 −
2RT

Ea
(12) maximum rate of weight loss and peak degradation temperature. (Rm)
T2 βEa Ea RT can be calculated using Eq. (19) as given below (Saeed et al., 2020a;
For n = 1, f(α) = (1 − α) and g(α) = − ln(1 − α) as shown in Eq. Ashraf et al., 2019b):
(13) (Ashraf et al., 2019a; Ouyang et al., 2022). DTGmax
[ ] [ ( )] Rm = (19)
− ln(1 − α) AR 2RT Ea Tp
ln = ln 1 − − (13)
T2 βEa Ea RT In Eq. (19) Tp represents peak degradation temperature while DTGmax
Hence, it can be written as presented in Eq. (14); represents the maximum rate of material weight loss and its unit is (wt
[ ] [ ] %/◦ C). In this study, a relationship is presented which is termed as py­
g(α) AR Ea rolysis factor (Pf) to evaluate the pyrolysis performance (Tariq et al.,
ln = ln − (14)
T2 βEa RT 2023; Munir et al., 2017) as shown in Eq. (20):
For measuring A, the following equation Eq. (15) can be used: DTGmax × DTGavg
Pf = (20)
Ts 2 × Tf
exp(c) βEa
A= (15)
R In Eq. (20) DTGmax and DTGavg represents the maximum and average
The intercept and its slope of the line are obtained by drawing a rate of materials weight loss and its unit is (wt%/◦ C), while Ts represents
[ ] the initial pyrolysis temperature and Tf represents the final pyrolysis
curve between ln − ln(1−T2α) and T1 is utilized for determining the (A)
temperature. This factor comprises the ease of devolatilization, volatile
and the (Ea). For n = 1, this plot turns out to be a straight line (Lu et al., ejection velocity, and the temperature of pyrolysis (Munir et al., 2017;
2013). The employed reaction models f(α) and g(α) utilized in the Saeed et al., 2020b).
Coats-Redfern integral technique are represented in Table 1 (Khan et al.,
2022; Khoja et al., 2023). The accurately determined data of the best
reaction mechanism of weight loss at a particular stage of C-PU blends
will be considered the most fitting model which is the objective for using
twelve g(α) with the highest R2 (Ganeshan et al., 2018; Mishra et al.,

911
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

the lesser will be the oxidation reactions thus releasing much energy
content (Dhar et al., 2022), and providing more hydrogen resources to
play a dominant role in hydrogenation (Zhang et al., 2021b). The H/C
atomic ratio of 100 C was (0.661) which is far less than (1.038) for
100PU. The (O/C) atomic ratio of 100 C was (0.278) and for 100PU was
(0.199).
The Gross calorific value (GCV) of C-PU blends determined is shown
in Fig. 2, which is in accordance with the carbon content in Table 2. It
defines the energy content in fuel and is considered an important se­
lection criterion for a sustainable conversion process (Channiwala and
Parikh, 2002). As listed in the literature, the value of 100 C’s GCV is
recorded as 30.23 MJ kg− 1 with a similar magnitude (Khan et al., 2022),
while for 100PU the value of GCV was 31.2 MJ kg− 1 analogous results
were stated in the literature for polyurethane (Stančin et al., 2021,
2022). The initial ignition temperature of 100PU is substantially less
than 100 C, and the burning speed is much faster than 100 C, and the PU
combustion process consists primarily of branch breaking, depolymer­
ization, fixed-carbon burning, and secondary polymer pyrolysis (Wang
Fig. 2. (C-PU) blends gross calorific value (GCV).
et al., 2020). The GCV of the three C-PU blends i.e. (75 C-25PU) was
30.04 MJ kg− 1, (50 C-50PU) was 30.61 MJ kg− 1, (25 C-75PU) was 30.84
3. Results and discussion MJ kg− 1. As the ratio of C in C-PU blends rises, a declining trend of GCV
from 100PU towards 100 C is observed from the evaluation of these
3.1. Characterization of C-PU blends results. The C-PU blends’ value sits in between the parent fuels’ values.
The sensible ratios of low-rank coal (25% and 50%) in the blends divulge
Table 2 represents the significant differences in the physiochemical enhancement in the general reactivity and ignition parameters also
properties of (100 C) and (100PU) through ultimate analysis. The indicating their suitable use for power generation.
analysis showed that the carbon, hydrogen, nitrogen, sulfur, and oxygen Fourier transform infrared spectroscopy (FTIR) was utilized to
content of 100 C was 68.91%, 3.82%, 1.28%, 0.48%, and 25.51% describe the vibration bands of C-PU blends. The vibration bands of
respectively, thus the sample coal lies in the category of bituminous coal. 100 C reveal the first absorption broad narrow spike at 3400 cm− 1 as
In the research literature, a similar CHNS-O composition for 100 C has presented in Fig. 3. This spike takes place from the absorption of hy­
been reported (Khoja et al., 2023; Coimbra et al., 2019). 100PU has droxyl stretching vibrational groups (–OH–) namely alcoholic or
nearly the same amount of carbon as of 100 C, but the other elemental phenolic compounds related to functional groups of oxygen, or minerals
contents are different. The carbon, hydrogen, nitrogen, sulfur, and ox­ like kaolin or halloysite, as well as carboxylic acid. As represented in
ygen content of 100PU was 69.30%, 6.04%, 6.29%, 0.00%, and 18.37% Fig. 3 at 2920 cm− 1 and 2850 cm− 1 in the region of (3000–2800 cm− 1),
respectively, similar CHNS-O composition has been reported in litera­ two stretching peaks can be seen corresponding to a group of methylene
ture (Wang et al., 2019; Stančin et al., 2022; Guo et al., 2021; Mikulčić (-CH2-) existing inside paraffin or a saturated aliphatic hydrocarbon.
et al., 2019). The presence of Oxygen content in 100 C and 100PU is Still, for 100 C, in the region of (1640–1520 cm− 1) at 1600 cm− 1, a
related to the amount of moisture that is present in it. The H/C atomic strong sharp peak was observed that originated through the aromatic
ratio evaluates the energy content of any fuel, the higher the H/C ratio compounds ring structure (-C– –C-) groups in the benzene ring signifying

Fig. 3. Fourier transform infrared spectroscopy (FTIR) of pure and C-PU blends.

912
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Fig. 4. TGA of C-PU blends (a) TG (b) enlarged first stage (c) enlarged second stage (d) enlarged third stage.

high carbon content in 100 C (Gohar et al., 2022). The narrow stretching because as 100PU and 100 C are mixed, the functional group’s intensity
absorption spike of 1440 cm− 1 originating in the region of changes. Fig. 3 demonstrates the peaks existing in blends (75 C-25PU),
(1520–1320 cm− 1) agrees with the vibrational stretching deformation of (50 C-50PU), and (25 C-75PU) discussed. The vibration bands existing
a methyl group (-CH3), which designates aliphatic or ring moieties and at 3313 cm− 1, 3324 cm− 1, and 3312 cm− 1 for the three C-PU blends
the broadening absorption spike of 1015 cm− 1 symbolizing (C-C) and respectively occur owing to the absorption groups of (–OH–) namely the
(C-H) having a plane deformation structures in aromatic. The narrowing compounds of alcoholic or phenolic-related functional groups of oxygen,
absorption spikes of 910 cm− 1 and 750 cm− 1 correspondingly origi­ and (N-H) stretching vibration modes. Another weak absorbance peak
nating in the region of (920–700 cm− 1) agree with the (=C-H) was identified at 2898 cm− 1, 2895 cm− 1, and 2892 cm− 1 for (75 C-
out-of-plane deformation vibrations in aromatic structures as presented 25PU), (50 C-50PU), and (25 C-75PU) respectively results from the
in Fig. 3 (Tauseef et al., 2022; Khan et al., 2022). (-CH2-) group situated inside a saturated aliphatic hydrocarbon. The
In the FTIR spectrum of 100PU the first strong absorption band was slightly strong stretching absorbance peak was observed at 1704 cm− 1,
found at 3309 cm− 1 that originates from the vibrational modes of (N-H) 1704 cm− 1, and 1700 cm− 1 for (75 C-25PU), (50 C-50PU), and (25 C-
stretching and another strong absorbance spike of 2974 cm− 1 corre­ 75PU) respectively caused by carbonyl compounds modes of (C– –O)
sponds to the asymmetrical stretching band of CH2 and was used as the stretching namely aldehydes, ketones, esters, or carboxyl aldehyde
normalized band. The sharp very strong stretching absorbance peak of changed with carbonyl influenced by conjugation (as with ketones) and
1711 cm− 1 agrees with the C– –O stretching modes of carbonyl com­ emerges in the blends when 100PU is blended with 100 C. In (75 C-
pounds such as ketones, aldehydes, esters, or carboxyl aldehyde changed 25PU), (50 C-50PU), and (25 C-75PU) respectively, a sharp absorption
with carbonyl influenced by conjugation (as with ketones). Another very spike originated in abundance at 1600 cm− 1, 1603 cm− 1, and
strong stretching absorbance band of 1595 cm− 1 corresponds to (C– –C) 1591 cm− 1 instigating from both the aromatic compounds ring structure
vinyl ether and an additional very strong spectrum of interest is iden­ (-C––C-) groups in the benzene ring vibrational stretching and (C– –C)
tified at 1509 cm− 1, which is instigated by the stretching modes of vinyl ether due to the blending of 100 C and 100PU at various ratios and
amide II in polyurethane and polyurea. A weak stretching absorption the peak intensity distribute equally among them also an additional very
peak was monitored at 1068 cm− 1, which is caused by (νa C-O-C) in the strong spectrum of interest is identified at 1513 cm− 1, 1505 cm− 1, and
ether while a strong absorption band was detected at 768 cm− 1 that 1505 cm− 1, which is instigated by stretching modes of amide II in
happens due to (C-C-O) in phase as presented in Fig. 3 (Wang et al., polyurethane and polyuria and emerges when 100PU is blended with
2013). 100 C. The slightly strong narrow absorption spikes of 1033 cm− 1,
The band intensity’s intermediate behavior is shown by C-PU blends 1037 cm− 1, and 1077 cm− 1 originating in (75 C-25PU), (50 C-50PU),

913
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Fig. 5. DTG of C-PU blends (a) DTG (b) enlarged first stage (c) enlarged second stage (d) enlarged third stage.

and (25 C-75PU) symbolizing (C-C) and (C-H) having a plane defor­ likewise, a slight WL% occurred due to moisture evaporation as shown
mation structures in aromatic along with an effect from (νa C-O-C) in in Fig. 4(b).
ether owing to blending of 100 C and 100PU, while a strong absorption A significant amount of WL% mainly occurs in the 2nd stage in which
band was detected at 752 cm− 1, 764 cm− 1, and 768 cm− 1 that occurs the degradation of C-PU blends takes place. In this stage, 100 C degra­
owing to (C-C-O) supporting the (=C-H) out-of-plane deformation vi­ dation starts at a higher temperature of 400 ◦ C and continues till 580 ◦ C,
brations in aromatic structures. This increasing trend from 100 C to­ mainly due to the high reactivity of coal, additionally, the 100 C has a
wards 100PU is understandable by the information that PU contains higher content of carbon, thus delaying the pyrolysis reaction and re­
many active NH and hydroxyl groups with high carbon and hydrogen sults described in the literature are also alike (Dwivedi et al., 2022).
content (Wang et al., 2020). While the pyrolysis of 100PU started earlier than 100C at a temperature
of 140 ◦ C and ends at around 400 ◦ C this earlier start of polymer
degradation is due to the weak linkages presence (Bhuvaneswari, 2018),
3.2. Thermo-Gravimetric Analysis (TGA) in N2 environment and containing pyrolytic degradation reactions of isocyanate and ure­
thane bonds (Jiang et al., 2018), the second WL% region occurs from
Fig. 4(a-d) and Fig. 5(a-d) represent the analysis of derivative ther­ around 420 ◦ C and goes on to 580 ◦ C where the remaining volatiles
mogravimetry (DTG) as (wt%/◦ C) and WL% of C-PU blends in regards to degraded in 100PU. 100PU degrades into melt (polyol) and gas when it
temperature (◦ C). The kinetic factors of co-pyrolysis and the rate of reaches the degradation temperature through the first sub-reaction and
degradation, WL%, are estimated with the help of TGA (Rasam et al., degrades further in the next stage (Li et al., 2016). For the three C-PU
2022). A three-stage thermal reaction was utilized for the character­ blends i.e. (75 C-25PU), (50 C-50PU), (25 C-75PU) in the second stage,
ization of the WL% behavior of C-PU blends (Yao et al., 2020b; Adewole the noteworthy WL% can be noticed designating the degradation of the
et al., 2020). The temperature ranges for the three stages are 25 –140 ◦ C, essential components of 100PU in the blends in the temperature region
140 – 580 ◦ C, and 580 – 900 ◦ C respectively. of 270 – 400 ◦ C, although some of the volatiles from 100PU degrades in
In the first stage of C-PU blends a slight WL% was observed which the temperature region of 410 – 580 ◦ C along with most of the WL% in
was due to the evaporation of moisture content, in this stage, no sig­ this range occurring due to the degradation of the constituents of 100 C
nificant peaks were observed for 100 C and 100PU. The slight WL% in in the C-PU blends as depicted in Fig. 4(c).
the first stage for 100PU was observed due to some low bowling point In the 3rd and final stage, slow degradation takes place from 580 –
matters, similar results were found in the literature (Xu et al., 2012). For 900 ◦ C for the C-PU blends. In this stage, the degradation of the solid
the three C-PU blends i.e. (75 C-25PU), (50 C-50PU), (25 C-75PU) in the residue of the 100 C occurs slowly and the WL% is small (Vhathvarothai
first stage, no significant difference due to blending happened, and

914
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

et al., 2014), and the reaction will take place with the development of Table 3
H2, CO, CO2, and char production (Lin et al., 2014). For 100PU the Synergistic effect calculations of (C-PU) blends.
degradation of the organic resins of the polyurethane foam in each stage Sample Name Experimental Calculated Deviation
(Jomaa et al., 2015), occurs slowly and the material residues degrade
WL% RL% WL% RL% WL% RL%
into small molecules and the WL% is slightly larger as compared to
100 C. Through the second sub-reaction, in 100PU, melt (polyol) further 100 C 27.10 72.90 - - - -
75 C-25PU 37.04 62.96 42.64 57.36 -13.13 9.76
degrades into solid residues and gas (polyol, H2CO, H2O, and CH4). For 50 C-50PU 66.60 33.40 58.19 41.81 14.45 -20.11
the three C-PU blends i.e. (75 C-25PU), (50 C-50PU), (25 C-75PU) in the 25 C-75PU 73.44 26.56 72.73 26.27 0.98 1.10
third stage, slow degradation of the solid residue that is mostly 100PU 89.32 10.68 - - - -
accredited to the naturally existing minerals and the carbonaceous res­
idues and takes place from 580 – 900 ◦ C as the parent fuels and the WL%
is slightly getting larger the PU’s ratio rises in the blends as shown in
Fig. 4(d).
A three-stage thermal reaction was utilized for the characterization
of C-PU blends’ DTG analysis. The temperature where the highest rate of
WL% happens throughout the procedure is called the peak degradation
temperature and it is designated as the Tp, while shown as a peak and the
maximum rate of WL% is designated as (wt%/◦ C) in the DTG curve. The
temperature ranges for the three stages are 25 –140 ◦ C, 140 – 580 ◦ C,
and 580 – 900 ◦ C respectively as displayed in Fig. 5(a-d).
In the 1st stage of C-PU blends there were no visible peaks as
demonstrated in Fig. 5(b). Whereas in the 2nd stage of C-PU blends
major peaks were witnessed separately for each blend. For 100 C in the
2nd stage reaction, the Tp is 435 ◦ C, and the (wt%/◦ C) is observed as
− 0.1523 wt%/◦ C. For 100 C, the peak DTG curve is ascribed to the
volatile matter release during the process of pyrolysis (Dwivedi et al.,
2019) as illustrated in Fig. 5(c). In the second stage for 100PU, the Tp is
330.71 ◦ C, and the (wt%/◦ C) is − 1.09 wt%/◦ C, while there was a nearly
flat second peak at 468.55 ◦ C and (wt%/◦ C) is − 0.1283 wt%/◦ C. The
main degradation stage is depicted by the 1st peak, in which many
broken urethane bonds, along with degraded polyols and isocyanates
are observed, and the end of the 1st peak is followed by a 2nd peak Fig. 6. Synergistic effects of C-PU blends.
which is nearly flat. The DTG curve 2nd peak could be termed 1,
1-Dichloro-1-fluoro ethane and analogous halogenated compounds’ after the reaction, the overall RL% was 72.9%, 62.96%, 33.40%,
secondary cracking along with soft segment’s degradation (Stančin 26.56%, and 10.68% for (100 C), (75 C-25PU), (50 C-50PU),
et al., 2021) as explained in Fig. 5(c). For the three C-PU blends i.e. (25 C-75PU), and (100PU) correspondingly.
(75 C-25PU), (50 C-50PU), (25 C-75PU) in the second stage two prom­
inent major peaks were observed as illustrated in Fig. 5(c). The Tp for 3.2.1. Synergistic effects
(75 C-25PU) was 329.16 ◦ C and (wt%/◦ C) was − 0.1716%/◦ C and 2nd Because of contrasts in their structural and chemical properties,
peak’s temperature was 447.43 ◦ C and (wt%/◦ C) was − 0.1374 wt 100 C and 100PU showed different thermal behavior. Thermal degra­
%/◦ C. For (50 C-50PU) the Tp was 320.90 ◦ C and the (wt%/◦ C) was dation of C-PU blends was studied in TGA to investigate their synergistic
− 0.6095 wt%/◦ C while the second peak was at 442.40 ◦ C and (wt effects during co-pyrolysis. The theoretical and experimental TG curves
%/◦ C) was − 0.1704 wt%/◦ C. The Tp for (25 C-75PU) was 330.24 ◦ C along with the deviation (%) in the results of the C-PU blends as shown
and the (wt%/◦ C) was − 0.7621 wt%/◦ C and the 2nd peak was at in Table 3 and are plotted in Fig. 6. The solid and the dotted lines
448.34 ◦ C and the (wt%/◦ C) was − 0.1375 wt%/◦ C. The C-PU blends represent the experimental and calculated values, respectively.
follow the same pattern as their parent fuels as the PU’s ratio rises in the The synergistic effects’ importance cannot be disregarded in
blends the first peak gets intensified and represents the degradation of improving the properties of pyrolytic oil and existing pyrolysis pro­
urethane bonds and isocyanates and polyols the end of the 1st peak is cesses. The synergistic effect is also contributed by heat transfer
immediately followed by another small peak. The DTG curve’s 2nd peak throughout co-pyrolysis (Merdun and Laougé, 2021). The combined
is described as a cracking of the carbon compounds present in C along reaction of volatile-volatiles and volatile-char is depicted by the differ­
with the degradation of the remaining volatiles of PU. The peak intensity ence in calculated and experimental values (Zhang et al., 2022b). The
of the first peak increases gradually as PU’s ratio in the C-PU blends rises coke formation is decreased and the yield of aromatics rises due to the
whereas the second peak intensity also increases but in an irregular introduction of hydrogen-rich plastic waste (100PU) as a co-reactant in
pattern. As the Tp decreases gradually in both peaks showing the co-pyrolysis with coal. The H/C ratio rises as shown in Table 2 due to
dominance of reactivity of PU in the blends. In the third stage of C-PU plastic possessing a higher hydrogen content than coal and the hydrogen
blends there were no obvious peaks observed due to the slow degrada­ donor/acceptor mechanism may affect the coal-plastic co-pyrolysis
tion of the blends as illustrated in Fig. 5(d). mechanism and thus modifying the reaction mechanism for oxygen’s
100 C’s Tp was larger than 100PU’s Tp. The fuel with the lowest Tp is elimination by replacing decarboxylation and decarbonylation reactions
the most reactive and has the highest WL% rate. Depending on the with dehydration reactions (Alam et al., 2020). The insight into radicals
blending ratio, the functioning of the essential materials throughout the reaction, exchanges, and hydrogen donors is given by the presence of
co-pyrolysis is mirrored by the WL% behavior of the blends. The low synergistic effects inside coal-plastic blends which are very complex at
volatile content of 100 C had an influence on the WL% of the blends, high temperatures (Hong et al., 2022; Stančin et al., 2022). Interactions
thereby increasing the quantity of C in the blends, and the WL% of C-PU between volatiles, volatiles and feedstock, and volatiles and char have
blends was reduced (Khan et al., 2022). After the reaction, the overall all contributed to the observed synergistic effect during the co-pyrolysis
WL% was 27.1%, 37.04%, 66.6%, 73.44%, and 89.32% for (100 C), process, which can dramatically impact the products (Zhang et al.,
(75 C-25PU), (50 C-50PU), (25 C-75PU), and (100PU) respectively. Also

915
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

2023). Co-pyrolysis of coal with plastic is a radical-driven process with


several reaction pathways of degradation, polymerization, recombina­
tion, hydrogenation, and secondary reactions occurring concurrently
(Feng et al., 2023). Because of the differences in the characteristics of
plastics, co-pyrolysis of different types of plastics with coal produces
diversified effects (Zhang et al., 2023).
In Fig. 5 for 75 C-25PU, a small variation between the calculated and
experimental values may be observed from the TG curve in the 1st stage
thus indicating a certain amount of positive synergy between 100 C and
100PU at an earlier lower temperature. The variation amid the theo­
retical and experimental values observed in the 2nd stage which is the
main pyrolysis zone depicts the presence of positive synergy at first but
suddenly changes from positive to negative at around 380 ◦ C between
100 C and 100PU. In the third stage, a huge gap amid the theoretical and
experimental values was observed and the synergy remained negative
towards the end at a higher temperature. WL% deviation values in
percentage were − 13.13%, for 75 C-25PU although the positive syn­
ergistic effects of 9.76% in relation to RL% deviance were observed that
gives the implication of synergistic effects’ minute role, importance and
suitability for the manufacturing of char as shown in Table 3. For 50 C-
50PU in all three stages, the synergy between 100 C and 100PU remains
positive as there was a significant difference amid the theoretical and the
experimental values from the start. The percentage WL% deviance was
14.45% as the positive synergistic effects, while RL% deviance of
− 20.11% was observed for the formation of pyrolytic oil as shown in
Table 3. For 25 C-75PU, the 1st stage of the TG curve shows that there
was a large gap amid the theoretical and the experimental values thus
indicating a certain amount of positive synergy between 100 C and
100PU at an earlier lower temperature. The gap between the calculated
and experimental values observed in the 2nd stage, which is the main
pyrolysis zone, depicts the presence of positive synergy at first but
suddenly changes from positive to negative at around 400 ◦ C between
100 C and 100PU. In the 3rd stage there was a minor deviance amid the
calculated and experimental values and the synergy remained negative
towards the end at a higher temperature. WL% deviance for 25 C-75PU
was 0.98%, while RL% deviance of 1.10% was observed, thus deemed as
suitable for both char and oil formation because of positive synergistic
effects as shown in Table 3. Positive synergistic effects can be seen in
terms of WL% for 50 C-50PU and 25 C-75PU which gives the implication
Fig. 7. Activation energy (Ea) for various models of C-PU blends (a) First Stage
of probable synergistic effects that influenced the manufacturing of gas
(b) Second Stage.
and oil yield due to volatiles’ release, 25 C-75PU also showed suitability
in char formation.
PU blends in TGA as presented in Table. S1. Based on R2 the Ea and A
It is thus concluded that due to the blending of 100 C and 100PU
were calculated for the selected twelve reaction mechanisms as
significant synergistic effects were observed at some points when they [ ]
are co-pyrolyzed, which leads to a lower temperature for the C-PU mentioned in Table 1, they were examined and graphed between ln g(a)
T2
blends, meaning less energy is consumed for co-pyrolysis. WL% and RL and T1 at a constant β of 10 ◦ C/min. The activation energy (Ea) was
% measured theoretically are less in agreement as compared to their calculated from the slope of the plot. The minimal amount of energy
values for a mixture with greater plastic content. WL% deviance values necessary to commence the reaction is represented by Ea, which also
have higher deviation for both 50 C-50PU and 25 C-75PU while in terms determines its reactivity for the samples (Naqvi et al., 2020), whereas A
of RL%, it is higher for 75 C-25PU. Plastic material is a more dominant represents the frequency of engaging molecules in a direction to
compound than coal and its degradation mechanism is simpler as commence the reaction and the material structure is more related to the
compared to coal. The overall synergistic effect has been deconvoluted A (Adewole et al., 2020). The precision of fitting data for R2 has a
into different types of interaction based on their source and assessed specified range from 0.90 to 0.99 (Hameed et al., 2018).
how they contribute to the total synergistic effect during co-pyrolysis For the parent fuels i.e. 100 C and 100PU with individual Ea, A, and
(Zhang et al., 2023). A strong synergistic effect between coal and plas­ R2, the active pyrolysis region is considered a single-stage process and is
tic is accomplished for plastic content (≤50%) in the blending ratio the rapid degradation zone that lies in the temperature region of 220 –
which means that level of synergy decreases with the plastic content 581 ◦ C and is used to perform the kinetic study due to the greatest WL%
increment. Therefore, 50 C-50PU is deemed as the optimum blend as percentage in this zone (Naqvi et al., 2019), while for the three blends i.
established by the synergistic effects. e. 75 C-25PU, 50 C-50PU, and 25 C-75PU the active pyrolysis region is
distributed into two stages, the temperature region for the first stage in
3.3. Thermo-Kinetic analysis this region is between 220 – 405 ◦ C and for the second stage in this
region is between 383 – 581 ◦ C with an individual Ea, A, and R2 for each
The coats-Redfern integral technique was utilized to have a detailed stage to perform the kinetic study.
insight into the kinetic profile and evaluate the effect of the blending The 100PU has the highest Ea and A than 100 C for every reaction
ratio on co-pyrolytic kinetic parameters. The kinetic triplet i.e. (Ea), (A), mechanism for the single stage. The kinetic results of (C-PU) blends are
and (R2) were measured according to the results of the co-pyrolysis of C-

916
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

displayed in Fig. 7 and the Table. S1. For 100 C, F2 with 0.979 as
maximum R2 value was the best model with an A of 1479.6 min− 1 and Ea
of 60.08 kJ/mol, while for 100PU, F3 with 0.979 as maximum R2 value
was the best model with an A of 25.25×107 min− 1 and Ea of 100.8 kJ/
mol. So these models were the most probable mechanism of 100 C and
100PU thermal degradation as shown in Fig. 7(a). The F2 and F3
chemical reaction order models for the parent fuels 100 C and 100PU
suggest that a basic reaction order defines the chemical reaction that is
the rate-controlling step which is the best descriptor of the theoretical
pyrolysis process commonly happening at a reaction interface in the
solid-state reactions. These models revealed that individual particle
degradation is accomplished through random nucleation of two and
three nuclei (Liu et al., 2015; Jiang et al., 2018).
For the three C-PU blends i.e. (75 C-25PU), (50 C-50PU), and (25 C-
75PU) in the first stage of the active pyrolysis region that occurs in the
temperature region of 220 – 405 ◦ C. For (75 C-25PU), the D4 model with
0.982 as the maximum R2 value was the best model with an A of
222776.12 min− 1 and Ea of 92.68 kJ/mol. For 50 C-50PU, the F2 model
with 0.976 as the maximum R2 value was the best model with an A of
75232.23 min− 1 and Ea of 67.01 kJ/mol. For (25 C-75PU), D3 model
with 0.969 as maximum R2 value was the best model with an A of
17.33×106 min− 1 and Ea of 108.29 kJ/mol so these were the most
possible mechanism of (75 C − 25PU), (50 C-50PU), and (25 C-75PU)
thermal degradation as shown in Fig. 7(a). In solid-state reaction which
follows the D4 reaction mechanism also known as the Ginstling
Brounstein equation is a function that depicts a diffusion-controlled
process that commences on the outermost surface of a spherical parti­
cle for the (75 C-25PU) blend (Alshehri et al., 2000). The F2 chemical
reaction order mechanism is the rate-determining step having a closer
description of the theoretical co-pyrolysis process generally occurring at
a reaction interface revealed that individual particle degradation is
accomplished through random nucleation of two nuclei for the
(50 C-50PU) blend (Liu et al., 2015). D3, a reaction mechanism known
as the Jander equation or three-way transport is employed in
three-dimensional diffusion for diffusion-controlled solid-state reac­
tivity kinetics in a sphere. Here diffusion in all three directions is
important for the (25 C-75PU) blend (Yorulmaz and Atimtay, 2009; Gil
et al., 2010).
The 2nd stage of the active pyrolysis region lies in the temperature
Fig. 8. Change in enthalpy (ΔH) for various models of C-PU blends (a) First
region of 383 – 581 ◦ C. In this stage from 383.6◦ to 580.7 ◦ C for the
Stage (b) Second Stage.
(75 C-25PU) blend, the F2 model had 0.99 as the maximum R2 value
with an A of 84.27 min− 1 and Ea of 39.09 kJ/mol. The D3 model had
be more complicated as the value of A gets larger (Ming et al., 2020).
0.989 as the maximum R2 value with an A of 0.71 min− 1 and Ea of
When A is less than 109 min− 1, it indicates a surface reaction; when A is
26.74 kJ/mol for the (50 C-50PU) blend. While D3 model with 0.993 as
greater than 109 min− 1, it indicates a simpler activated complex (Hu
the maximum R2 value was the best model with an A of 0.28 min− 1 and
et al., 2019). This shows that the pyrolysis of 100 C has influenced the
Ea of 20.78 kJ/mol for the (25 C-75PU) so these were the most possible
structure of 100PU, somewhat reducing the Ea of 100PU in the C-PU
mechanisms of (75 C − 25PU), (50 C-50PU), and (25 C-75PU) thermal
blends. Also, the 100PU degradation mechanism is more like biomass
degradation in the 2nd stage as depicted in Fig. 7(b). In solid-state re­
than plastics as the polyurethane foam Ea is lower than the rest of the
action for the (75 C-25PU) blend the F2 chemical reaction order
plastics (Stančin et al., 2021). The Ea and the A were further reduced for
mechanism is the rate-determining step having a closer description of
both the pure samples of 100 C and 100PU through the best blend of
the theoretical co-pyrolysis process generally occurring at a reaction
50 C-50PU, and hence we agreed with its selection in terms of Ea for
interface revealed that individual particle degradation is accomplished
co-pyrolysis reactions.
through random nucleation of two nuclei (Jiang et al., 2018). The D3
Engineering tools were developed for the co-pyrolysis of C-PU blends
reaction mechanism is also known as the jander equation, whilst
through thermodynamic characteristics i.e. (ΔS), (ΔH), and (ΔG) were
three-dimensional diffusion is employed for diffusion-controlled solid-­
obtained via Tp through DTG analysis for the two stages as demonstrated
state processes in a sphere for the (50 C-50PU) and (25 C-75PU) blend,
in Table. S2. These parameters were obtained from single-stage pro­
in diffusion-controlled reactions, several chemical reactions or
cesses with an individual (ΔH), (ΔG), and (ΔS) in 100 C and 100PU,
micro-structural changes in solids take place which indicates the
while for C-PU blends it is obtained from two-stage processes with an
movement and transportation of gas molecules in the solid phase (Liu
individual (ΔH), (ΔG), and (ΔS) for the twelve reaction mechanisms.
et al., 2015; Gil et al., 2010).
These parameters are appreciated in gaining an insight into the reac­
The Ea for C-PU blends varies a lot in all the twelve reaction mech­
tivity, the direction of energy conversion, and the feasibility of the re­
anisms. As the ratio of PU increases, in some reaction mechanisms the Ea
action as presented in Fig. 8(a,b) and Fig. 9(a,b) (Balogun).
varies i.e. it decreases and increases in the 1st stage, whereas in the 2nd
For both the parent fuels (ΔS) have negative, while (ΔH) and (ΔG)
stage, the Ea varies in an increasing and decreasing manner for the C-PU
had positive values for the twelve reaction mechanisms in the first stage.
blends. In both the first and second stages, the pre-exponential factor
100 C has a temperature region of 340.6 – 580.8 ◦ C, model F2 was the
value increases as PU’s ratio increases in C-PU blends. The reaction will

917
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Table 4
Thermo-Kinetic physiognomies of C-PU blends co-pyrolysis after selection of the
best model.
Parameters 100 C 75 C- 50 C- 25 C-75PU 100PS
25PU 50PU

Section I
G(α) F2 D4 F2 D3 F3
Ea (kJ.mol− 1) 60.08 92.68 67.01 108.29 100.80
A (min− 1) 1479.6 2.23 × 7.5 × 104 17.33 × 25.25 ×
105 106 107
R2 0.979 0.982 0.976 0.969 0.979
ΔH (kJ.mol− 1) 54.06 87.66 61.99 103.26 95.78
ΔG (kJ.mol− 1) 198.75 182.30 162.10 176.12 155.05
ΔS (kJ.mol− 1. -0.19 -0.16 -0.17 -0.12 -0.09
K− 1)
Section II
G(α) - D3 D3 D3 -
Ea (kJ.mol− 1) - 39.09 26.74 20.78 -
A (min− 1) - 84.27 0.71 0.28 -
R2 - 0.999 0.989 0.993 -
ΔH (kJ.mol− 1) - 33.07 20.70 14.75 -
ΔG (kJ.mol− 1) - 195.15 212.11 211.54 -
ΔS (kJ.mol− 1. - -0.22 -0.26 -0.27 -
K− 1)

75PU) has an (ΔS), (ΔH), and (ΔG) of − 0.27 kJ/mol.k, 14.75 kJ/mol,
and 211.54 kJ/mol as illustrated in Fig. 8(b) and Fig. 9(b).
The behavior of the reaction mechanism is depicted by change in
enthalpy (ΔH) as an endothermic and as an exothermic. The entire C-PU
blends co-pyrolysis process is considered endothermic as the values of
(ΔH) are positive (Badshah et al., 2021a). (ΔH) was higher for 100PU
than that of 100 C. A higher value of (ΔH) designates a higher amount of
energy is essential for the degradation process to disrupt the bonds of the
reactant sample (Parthasarathy et al., 2021). For the parent fuels, (5.02 –
6.02) kJ/mol was the variation range between Ea and ΔH in the single
stage and the variation range in the two stages for the three blends was
between (5.0 – 6.05) kJ/mol which indicated that the reaction probably
proceeded due to the small potential energy barrier (Badshah et al.,
2021b; Florentino-Madiedo et al., 2021). Information about the overall
increase in energy of the system is provided by (ΔG) as the activated
complexes are created and the reactants are consumed. For 100 C and
Fig. 9. Change in Gibbs Free Energy (ΔG) for various models of C-PU blends (a) 100PU, during the whole process, (ΔG) has always a positive value in
First Stage (b) Second Stage. the single stage, and in the two stages for the (C-PU) blends indicating
that the reaction requires external heat and is non-spontaneous that
best that has an (ΔS), (ΔH), and (ΔG) of − 0.19 kJ/mol.k, 54.06 kJ/mol, indicates 100 C is more non-spontaneous than 100PS (Vasudev et al.,
and 198.75 kJ/mol whereas 100PU has a temperature region of 240 – 2020; Hu et al., 2022). The low (ΔS) indicates that while the solid ma­
416.5 ◦ C, model F3 was the best having an (ΔS), (ΔH), and (ΔG) of terial is in thermodynamic equilibrium condition it was subjected to a
− 0.09 kJ/mol.k, 95.78 kJ/mol, and 155.05 kJ/mol, as shown in Fig. 8 variety of physical and chemical processes (Saeed et al., 2021b). For all
(a) and Fig. 9(a). For the three (C-PU) blends i.e. (75 C-25PU), (50 C- the stages of the C-PU blends during the whole process, (ΔS) has nega­
50PU), and (25 C-75PU) in the first stage of the main pyrolysis region tive values. The activated complex of C-PU blends when compared to the
that happens in the temperature region of 220 – 405 ◦ C. (ΔS) have preliminary constituent, had a more ordered structure, and a
negative values while (ΔH) and (ΔG) had positive values for the twelve well-ordered structure from the chaotic structure was involved in the
reaction mechanisms. For the (75 C-25PU) blend the D4 model has an co-pyrolysis. The lower value of ΔS indicates lower activity and takes
(ΔS), (ΔH), and (ΔG) of − 0.16 kJ/mol.k, 87.66 kJ/mol, and 182.3 kJ/ more time to move towards thermodynamic equilibrium and will reach
mol. The model F2 was best for (50 C-50PU) with an (ΔS), (ΔH), and their soon (Parthasarathy et al., 2021).
(ΔG) of − 0.17 kJ/mol.k, 61.99 kJ/mol, and 162.10 kJ/mol. While the The best blend (50 C-50PU) had the lowermost values of ΔH and ΔG,
best model was D3 for (25 C-75PU) with an (ΔS), (ΔH), and (ΔG) of decreasing it for 100 C and 100PU, lower ΔG value means that it ne­
− 0.12 kJ/mol.k, 103.26 kJ/mol, and 176.12 kJ/mol as demonstrated cessitates a lesser amount of energy for activated complex creation,
in Fig. 8(a) and Fig. 9(a). whereas a higher ΔG value will represent lower reaction favorability,
The (ΔH), (ΔG), and (ΔS) values for both parent fuels weren’t thus making it more non-spontaneous as well as ΔS has a higher negative
computed in the 2nd stage. The second stage for the three C-PU blends i. value for the best blend, indicating that the reaction in the best blend is
e. (75 C-25PU), (50 C-50PU), and (25 C-75PU) takes place in the tem­ approaching equilibrium. The thermo-kinetic physiognomies of C-PU
perature region of 383 – 581 ◦ C. The (ΔS) values were negative whereas blend co-pyrolysis after the selection of the best model as shown in
(ΔH) and (ΔG) has positive values for the twelve reaction mechanisms. Table 4. According to the thermo-kinetic model-fitted study, C-PU
For the (75 C-25PU) blend the F2 model has an (ΔS), (ΔH), and (ΔG) of blends are suitable for boosting the production of sustainable fuels and
− 0.22 kJ/mol.k, 33.07 kJ/mol, and 195.15 kJ/mol. The best model D3 successfully turning them into useful products.
model for (50 C-50PU) has an (ΔS), (ΔH), and (ΔG) of − 0.26 kJ/mol.k,
20.70 kJ/mol, and 212.11 kJ/mol. While the best model D3 for (25 C-

918
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Table 5
Reactivity Analysis of 100 C, 100PU, and 50 C-50PU.
3
Sample Name Ts (◦ C) Tp (◦ C) Tf (◦ C) DTGmax (wt%/◦ C) DTGavg (wt%/◦ C) Rm (10− wt%/min. ◦ C) Pf (10− 8)

100 C 320 435 560 -0.1523 -0.6421 0.3501 × 10 − 3


0.705 × 10− 8

50 C-50PU 230 320.9 380 -0.6095 -2.4750 1.899 × 10− 3 7.503 × 10− 8

100PU 240 330.7 406 -1.09 -4.0434 3.296 × 10− 3 18.85 × 10− 8

4. Conclusions

In this work, co-pyrolysis of low-rank coal-polyurethane foam waste


C-PU blends has been investigated to outperform the pyrolysis of indi­
vidual materials through reactivity and thermo-kinetic analysis for the
production of pyrolytic oil, gas, and char. The CHNS-O and GCV values
of parent fuels were nearly equal to each other indicating their suitable
use for power generation. The C-PU blends were pyrolyzed in TGA thus
showing their degradation mechanism concerning the profile of the TG
curve generally associated. The TGA analysis indicated that the thermal
degradation behavior could be distributed into three stages having
temperature regions viz. (25 –140 ◦ C), (140 – 580 ◦ C), and (580 –
900 ◦ C) respectively. Synergistic effects of the C-PU blends were tracked
using the data obtained from TGA analysis and calculated data, using the
deviation method. The blend (50 C-50PU) had the highest WL% value of
positive synergistic effects and was chosen as the best blend in the
production of pyrolytic oil. The model-fitting kinetics method i.e. the
coats-Redfern was utilized to investigate via the twelve reaction mech­
Fig. 10. Mean reactivity and pyrolysis factor of 100 C, 100PU, and 50 C-50PU. anistic models. The apparent kinetics were obtained by selecting the best
model based on the highest R2 value. For 100 C and 100PU the active
3.4. Reactivity analysis pyrolysis region is considered a single-stage process while the three
blends i.e. 75 C-25PU, 50 C-50PU, and 25 C-75PU divided into two
The reactivity of low-rank coal (100 C), polyurethane plastic stages to perform the thermo-kinetic study. In the first degradation
(100PU), and the optimum blend (50 C-50PU) materials during the stage, the F3 and F2 models were the best for 100PU and 100 C having
degradation process has been investigated using two methods which are (Ea) of 100.8 kJ/mol and 60.08 kJ/mol respectively. While the optimum
mean reactivity (Rm) and pyrolysis factor (Pf). The method established blend (50 C-50PU), F2 and D3 models were optimal having an (Ea) of
by (Ghetti et al., 1996) (Ghetti et al., 1996), was obtained to calculate 67.01 and 26.74 kJ/mol for both stages, so it lowered the (Ea) of 100PU
the Relative mean reactivity (Rm). According to the values of Rm given in in both stages, proving its selection and feasibility, stable product for­
Table 5 and Fig. 10 the rank of the sample can be ordered as mation, and spontaneity were confirmed through thermodynamic
100PU> 50 C-50PU> 100 C in an inert environment forming a curve characterizations. (ΔS), (ΔH), and (ΔG) in the 1st stage was − 0.17 kJ/
like a convex mirror shape. This means higher the volatile content, the mol.k, 61.99 kJ/mol, and 162.10 kJ/mol, while, for the 2nd, they were
reactivity and the ignition of the samples are higher and easier (Munir − 0.26 kJ/mol.k, 20.70 kJ/mol, and 212.11 kJ/mol and showed a more
et al., 2017). The total WL% was 27.1%, 89% and 66.6% for 100 C, systematized structure as related to parent fuels. The performance
100PU, and 50 C-50PU, and yielded Rm having a value of (0.3501 × analysis in terms of Rm and Pf showed that 100PU and 50 C-50PU were
10− 3 ), (1.899 × 10− 3 ), and (3.296 × 10− 3 ). The pyrolytic performance more reactive than 100 C which has the lowest performance values and
of fuel can be estimated efficiently by calculating the pyrolysis factor low devolatilization rates. These findings specify the probability of C-PU
(Pf). Pf compares pyrolytic performance by combining devolatilization blend co-pyrolysis for energy recovery as well as an effective waste
speed and pyrolysis temperatures. It quantitatively describes the devo­ plastics treatment technology. They may contribute to the construction
latilization behavior (Ashraf et al., 2019b). It is evident from the results of a commercially viable co-pyrolysis reactor system, but further sus­
reported in Table 5 and Fig. 10 that Pf value follows an increasing trend tainability and cost-benefit analyses are needed to make it more feasible.
from 100 C towards 100PU forming a curve like a concave mirror shape. The addition of a suitable catalyst will enhance reactivity and reduce the
The values of Pf showed that it was higher for more reactive fuels like energy requirement.
100PU is more reactive than 100 C and has better pyrolysis performance
than it which would be due to the higher devolatilization rates of plastic Declaration of Competing Interest
compared to coal. The optimum blend (50 C-50PU) showed optimum
pyrolysis performance when 100 C is blended with 100PU. The value of The authors declare that they have no known competing financial
Pf are (0.705 × 10− 8 ), (18.85 × 10− 8 ), and (7.503 × 10− 8 ) for 100 C, interests or personal relationships that could have appeared to influence
100PU, and 50 C-50PU respectively which means that the pyrolysis the work reported in this paper.
factor is improved by blending 100PU with 100 C. This could be owing
to 100PU’s extremely high hydrogen bond basicity due to high hydrogen Acknowledgments
content as compared to 100 C, which renders it highly likely to disrupt
hydrogen bond formation in the 100 C structure (Saeed et al., 2020a). The authors extend their appreciation to the Deanship of Scientific
The fuel samples followed the same order for pyrolysis factor (Pf) as for Research at King Khalid University for funding this work through the
the mean reactivity (Rm). This analysis further confirms our selection of Large Group Research Project under grant number RGP2/258/44)
the optimum blend 50 C-50PU for the pyrolysis process accordingly with
kinetics and thermodynamics study. Authors’ contribution statement

Z.M.B and W.H.K: Investigation, Data Curation, and Writing-

919
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Original draft. Gil, M.V., Casal, D., Pevida, C., Pis, J.J., Rubiera, F., 2010. Thermal behaviour and
kinetics of coal/biomass blends during co-combustion. Bioresour. Technol. 101,
M.A, M.H and A.H.K: Conceptualization, Methodology, Writing –
5601–5608.
original draft, Supervision. Gohar, H., Khoja, A.H., Ansari, A.A., Naqvi, S.R., Liaquat, R., Hassan, M., Hasni, K.,
H.A.G, Z.M.E.B, and R.L: Resources, Funding acquisition, and Qazi, U.Y., Ali, I., 2022. Investigating the characterisation, kinetic mechanism, and
Reviewing-editing Original draft. thermodynamic behaviour of coal-biomass blends in co-pyrolysis process. Process
Saf. Environ. Prot. 163, 645–658.
Gouws, S.M., Carrier, M., Bunt, J.R., Neomagus, H.W.J.P., 2021. Co-pyrolysis of coal and
Appendix A. Supporting information raw/torrefied biomass: a review on chemistry, kinetics and implementation. Renew.
Sustain. Energy Rev. 135, 110189.
Guo, X., Li, N., Zhang, T., 2021. Preparation of hydrogen-rich gas from waste
Supplementary data associated with this article can be found in the polyurethane foam by steam gasification and catalytic reforming in a two-stage fixed
online version at doi:10.1016/j.psep.2024.01.041. bed reactor. J. Mater. Cycles Waste Manag. 23, 1955–1963.
Hameed, Z., Aman, Z., Naqvi, S.R., Tariq, R., Ali, I., Makki, A.A., 2018. Kinetic and
thermodynamic analyses of sugar cane bagasse and sewage sludge co-pyrolysis
References process. Energy Fuels 32, 9551–9558.
Hassan, H., Lim, J.K., Hameed, B.H., 2016. Recent progress on biomass co-pyrolysis
conversion into high-quality bio-oil. Bioresour. Technol. 221, 645–655.
Adewole, B.Z., Adeboye, B.S., Malomo, B.O., Obayopo, S.O., Mamuru, S.A., Asere, A.A.,
Hong, D., Gao, P., Wang, C., 2022. A comprehensive understanding of the synergistic
2020. CO-pyrolysis of bituminous coal and coconut shell blends via
effect during co-pyrolysis of polyvinyl chloride (PVC) and coal. Energy 239, 122258.
thermogravimetric analysis. Energy Sources, Part A: Recovery, Util., Environ. Eff.
Hu, J., Yan, Y., Evrendilek, F., Buyukada, M., Liu, J., 2019. Combustion behaviors of
1–14.
three bamboo residues: gas emission, kinetic, reaction mechanism and optimization
Alam, M., Bhavanam, A., Jana, A., Viroja, Jk.S., Peela, N.R., 2020. Co-pyrolysis of
patterns. J. Clean. Prod. 235, 549–561.
bamboo sawdust and plastic: synergistic effects and kinetics. Renew. Energy 149,
Hu, Y., Liu, J., Li, X., Wang, F., Luo, L., Pei, Y., Lei, Y., Tang, K., 2022. Assessment of the
1133–1145.
pyrolysis kinetics and mechanism of vegetable-tanned leathers. J. Anal. Appl.
Ali, I., Naqvi, S.R., Bahadar, A., 2018. Kinetic analysis of Botryococcus braunii pyrolysis
Pyrolysis 164, 105502.
using model-free and model fitting methods. Fuel 214, 369–380.
Jayanti, S., Saravanan, V., Sivaji, S., 2012. Assessment of retrofitting possibility of an
Alshehri, S.M., Monshi, M.A.S., Abd El-Salam, N.M., Mahfouz, R.M., 2000. Kinetics of the
Indian pulverized coal boiler for operation with Indian coals in oxy-coal combustion
thermal decomposition of γ-irradiated cobaltous acetate. Thermochim. Acta 363,
mode with CO2 sequestration. Proc. Inst. Mech. Eng., Part A: J. Power Energy 226,
61–70.
1003–1013.
Ashraf, A., Sattar, H., Munir, S., 2019a. A comparative applicability study of model-
Jiang, L., Zhang, D., Li, M., He, J.-J., Gao, Z.-H., Zhou, Y., Sun, J.-H., 2018. Pyrolytic
fitting and model-free kinetic analysis approaches to non-isothermal pyrolysis of coal
behavior of waste extruded polystyrene and rigid polyurethane by multi kinetics
and agricultural residues. Fuel 240, 326–333.
methods and Py-GC/MS. Fuel 222, 11–20.
Ashraf, A., Sattar, H., Munir, S., 2019b. Thermal decomposition study and pyrolysis
Jin, Q., Wang, X., Li, S., Mikulčić, H., Bešenić, T., Deng, S., Vujanović, M., Tan, H.,
kinetics of coal and agricultural residues under non-isothermal conditions. Fuel 235,
Kumfer, B.M., 2019. Synergistic effects during co-pyrolysis of biomass and plastic:
504–514.
Gas, tar, soot, char products and thermogravimetric study. J. Energy Inst. 92,
Badshah, S.L., Shah, Z., Francisco Alves, J.L., Gomes da Silva, J.C., Iqbal, A., 2021a.
108–117.
Pyrolysis of the freshwater macroalgae Spirogyra crassa: Evaluating its bioenergy
Jomaa, G., Goblet, P., Coquelet, C., Morlot, V., 2015. Kinetic modeling of polyurethane
potential using kinetic triplet and thermodynamic parameters. Renew. Energy 179,
pyrolysis using non-isothermal thermogravimetric analysis. Thermochim. Acta 612,
1169–1178.
10–18.
Badshah, S.L., Shah, Z., Alves, J.L.F., da Silva, J.C.G., Noreen, N., Iqbal, A., 2021b.
Khalid, U., Khoja, A.H., Daood, S.S., Khan, W.U.H., Din, I.U., Al-Anazi, A., Petrillo, A.,
Kinetic and thermodynamics study of the pyrolytic process of the freshwater
2024. Experimental and numerical techniques to evaluate coal/biomass fly ash blend
macroalga, Chara vulgaris. J. Appl. Phycol. 33, 2511–2521.
characteristics and potentials. Sci. Total Environ. 912, 169218.
A.O. Balogun, A.A. Adeleke, P.P. Ikubanni, S.O. Adegoke, A.M. Alayat, A.G. McDonald,
Khan, W.U.H., Khoja, A.H., Gohar, H., Naqvi, S.R., Din, I.U., Lumbers, B., Salem, M.A.,
Kinetics modeling, thermodynamics and thermal performance assessments of
Alzahrani, A.Y., 2022. In depth thermokinetic investigation on Co-pyrolysis of low-
pyrolytic decomposition of Moringa oleifera husk and Delonix regia pod.
rank coal and algae consortium blends over CeO2 loaded hydrotalcite (MgNiAl)
Bhuvaneswari, H., 2018. G, 3 - Degradability of Polymers. In: Thomas, S., Rane, A.V.,
catalyst. J. Environ. Chem. Eng. 10, 108293.
Kanny, K., V.K, A., Thomas, M.G. (Eds.), Recycling of Polyurethane Foams. William
Khoja, A.H., Gohar, H., Khan, W.U.H., Qazi, U.Y., Din, I.U., Al-Anazi, A., Ashraf, W.M.,
Andrew Publishing, pp. 29–44.
Mujtaba, M.A., Riaz, F., Daood, S.S., 2023. Exploring copyrolysis characteristics and
Cai, J., Wang, Y., Zhou, L., Huang, Q., 2008. Thermogravimetric analysis and kinetics of
thermokinetics of peach stone and bituminous coal blends. Energy Sci. Eng. 11,
coal/plastic blends during co-pyrolysis in nitrogen atmosphere. Fuel Process.
4302–4323.
Technol. 89, 21–27.
Li, K., Pau, D.S., Zhang, H., 2016. Pyrolysis of polyurethane foam: optimized search for
Channiwala, S.A., Parikh, P.P., 2002. A unified correlation for estimating HHV of solid,
kinetic properties via simultaneous K–K method, genetic algorithm and elemental
liquid and gaseous fuels. Fuel 81, 1051–1063.
analysis. Fire Mater. 40, 800–817.
Coimbra, R.N., Escapa, C., Otero, M., 2019. Comparative thermogravimetric assessment
Lin, Y., Li, Q., Ji, K., Li, X., Yu, Y., Zhang, H., Song, Y., Fu, Y., Sun, L., 2014.
on the combustion of coal, microalgae biomass and their blend. Energies 12, 2962.
Thermogravimetric analysis of pyrolysis kinetics of Shenmu bituminous coal,
Crescentini, T.M., May, J.C., McLean, J.A., Hercules, D.M., 2019. Mass spectrometry of
Reaction Kinetics. Mech. Catal. 113, 269–279.
polyurethanes. Polymer 181, 121624.
Liu, X., Chen, M., Wei, Y., 2015. Kinetics based on two-stage scheme for co-combustion
Dhar, S.A., Sakib, T.U., Hilary, L.N., 2022. Effects of pyrolysis temperature on production
of herbaceous biomass and bituminous coal. Fuel 143, 577–585.
and physicochemical characterization of biochar derived from coconut fiber biomass
Lu, K.-M., Lee, W.-J., Chen, W.-H., Lin, T.-C., 2013. Thermogravimetric analysis and
through slow pyrolysis process. Biomass-.-. Convers. Biorefinery 12, 2631–2647.
kinetics of co-pyrolysis of raw/torrefied wood and coal blends. Appl. Energy 105,
Ding, G., He, B., Yao, H., Kuang, Y., Song, J., Su, L., 2021. Synergistic effect, kinetic and
57–65.
thermodynamics parameters analyses of co-gasification of municipal solid waste and
D. Mallick, P. Mahanta, V.S. Moholkar, Synergistic Effects in Gasification of Coal/
bituminous coal with CO2. Waste Manag. 119, 342–355.
Biomass Blends: Analysis and Review, 2018.
Dwivedi, K.K., Shrivastav, P., Karmakar, M.K., Pramanick, A.K., Chatterjee, P.K., 2019.
Melendi-Espina, S., Alvarez, R., Diez, M.A., Casal, M.D., 2015. Coal and plastic waste co-
A comparative study on pyrolysis characteristics of bituminous coal and low-rank
pyrolysis by thermal analysis–mass spectrometry. Fuel Process. Technol. 137,
coal using thermogravimetric analysis (TGA). Int. J. Coal Prep. Util. 1–11.
351–358.
Dwivedi, K.K., Shrivastav, P., Karmakar, M.K., Pramanick, A.K., Chatterjee, P.K., 2022.
Merdun, H., Laougé, Z.B., 2021. Kinetic and thermodynamic analyses during co-pyrolysis
A comparative study on pyrolysis characteristics of bituminous coal and low-rank
of greenhouse wastes and coal by TGA. Renew. Energy 163, 453–464.
coal using thermogravimetric analysis (TGA). Int. J. Coal Prep. Util. 42, 1–11.
Mikulčić, H., Jin, Q., Stančin, H., Wang, X., Li, S., Tan, H., Duić, N., 2019.
Evode, N., Qamar, S.A., Bilal, M., Barceló, D., Iqbal, H.M.N., 2021. Plastic waste and its
Thermogravimetric analysis investigation of polyurethane plastic thermal properties
management strategies for environmental sustainability. Case Stud. Chem. Environ.
under different atmospheric conditions, journal of sustainable development of
Eng. 4, 100142.
energy. Water Environ. Syst.
Feng, W., Zheng, M., Jin, L., Bai, J., Kong, L., Li, H., Bai, Z., Li, W., 2023. Co-pyrolysis
Ming, X., Xu, F., Jiang, Y., Zong, P., Wang, B., Li, J., Qiao, Y., Tian, Y., 2020. Thermal
behaviors of coal and polyethylene by combining in-situ Py-TOF-MS and reactive
degradation of food waste by TG-FTIR and Py-GC/MS: Pyrolysis behaviors, products,
molecular dynamics. Fuel 331, 125802.
kinetic and thermodynamic analysis. J. Clean. Prod. 244, 118713.
Florentino-Madiedo, L., Vega, M.F., Díaz-Faes, E., Barriocanal, C., 2021. Evaluation of
Mishra, R.K., Sahoo, A., Mohanty, K., 2019. Pyrolysis kinetics and synergistic effect in co-
synergy during co-pyrolysis of torrefied sawdust, coal and paraffin. A kinetic and
pyrolysis of Samanea saman seeds and polyethylene terephthalate using
thermodynamic study. Fuel 292, 120305.
thermogravimetric analyser. Bioresour. Technol. 289, 121608.
Folkedahl, B.C., Snyder, A.C., Strege, J.R., Bjorgaard, S.J., 2011. Process development
Munir, S., Sattar, H., Nadeem, A., Azam, M., 2017. Thermal and kinetic performance
and demonstration of coal and biomass indirect liquefaction to synthetic iso-
analysis of corncobs, falsa sticks, and chamalang coal under oxidizing and inert
paraffinic kerosene. Fuel Process. Technol. 92, 1939–1945.
atmospheres. Energy Sources, Part A: Recovery, Util., Environ. Eff. 39, 775–782.
Ganeshan, G., Shadangi, K.P., Mohanty, K., 2018. Degradation kinetic study of pyrolysis
Naqvi, S.R., Hameed, Z., Tariq, R., Taqvi, S.A., Ali, I., Niazi, M.B.K., Noor, T., Hussain, A.,
and co-pyrolysis of biomass with polyethylene terephthalate (PET) using
Iqbal, N., Shahbaz, M., 2019. Synergistic effect on co-pyrolysis of rice husk and
Coats–Redfern method. J. Therm. Anal. Calorim. 131, 1803–1816.
sewage sludge by thermal behavior, kinetics, thermodynamic parameters and
Ghetti, P., Ricca, L., Angelini, L., 1996. Thermal analysis of biomass and corresponding
artificial neural network. Waste Manag. 85, 131–140.
pyrolysis products. Fuel 75, 565–573.

920
Z.M. Banyhani et al. Process Safety and Environmental Protection 184 (2024) 907–921

Naqvi, S.R., Ali, I., Nasir, S., Ali Ammar Taqvi, S., Atabani, A.E., Chen, W.-H., 2020. co-pyrolysis of waste rigid polyurethane foam and sawdust mixture. Renew. Energy
Assessment of agro-industrial residues for bioenergy potential by investigating 196, 1218–1228.
thermo-kinetic behavior in a slow pyrolysis process. Fuel 278, 118259. Tariq, R., Saeed, S., Riaz, M., Saeed, S., 2023. Kinetic and thermodynamic evaluation of
Nyoni, B., Duma, S., Bolo, L., Shabangu, S., Hlangothi, S.P., 2020. Co-pyrolysis of South almond shells pyrolytic behavior using Coats–redfern and pyrolysis product
African bituminous coal and Scenedesmus microalgae: Kinetics and synergistic distribution model. Energy Sources, Part A: Recovery, Util., Environ. Eff. 45,
effects study. Int. J. Coal Sci. Technol. 7, 807–815. 4446–4462.
Ouyang, S., Zhu, C., Xia, Y., Yang, Y., Zhang, C., Tsang, C.-W., Xiong, D., Li, L., Yang, K., Tauseef, M., Ansari, A.A., Khoja, A.H., Naqvi, S.R., Liaquat, R., Nimmo, W., Daood, S.S.,
2022. Synergistic effect of co-pyrolysis of tea seed shells and scrap tyres and product 2022. Thermokinetics synergistic effects on co-pyrolysis of coal and rice husk blends
evaluation. Biomass-.-. Convers. Biorefinery. for bioenergy production. Fuel 318, 123685.
Pan, R., Duque, J.V.F., Debenest, G., 2021. Investigating waste plastic pyrolysis kinetic Vasudev, V., Ku, X., Lin, J., 2020. Pyrolysis of algal biomass: determination of the kinetic
parameters by genetic algorithm coupled with thermogravimetric analysis. Waste triplet and thermodynamic analysis. Bioresour. Technol. 317, 124007.
Biomass-.-. Valoriz. 12, 2623–2637. Vhathvarothai, N., Ness, J., Yu, Q.J., 2014. An investigation of thermal behaviour of
Park, D.K., Kim, S.D., Lee, S.H., Lee, J.G., 2010. Co-pyrolysis characteristics of sawdust biomass and coal during copyrolysis using thermogravimetric analysis. Int. J. Energy
and coal blend in TGA and a fixed bed reactor. Bioresour. Technol. 101, 6151–6156. Res. 38, 1145–1154.
Parthasarathy, P., Fernandez, A., Al-Ansari, T., Mackey, H.R., Rodriguez, R., McKay, G., Wang, H., Wang, Q.-s, He, J.-j, Mao, Z.-l, Sun, J.-h, 2013. Study on the pyrolytic
2021. Thermal degradation characteristics and gasification kinetics of camel manure behaviors and kinetics of rigid polyurethane foams. Procedia Eng. 52, 377–385.
using thermogravimetric analysis. J. Environ. Manag. 287, 112345. Wang, H., Tian, Y., Chen, X., Li, J., 2020. Kinetic and synergetic effect analysis of the co-
Pinto, F., Costa, P., Gulyurtlu, I., Cabrita, I., 1999. Pyrolysis of plastic wastes. 1. Effect of combustion of coal blended with polyurethane materials. ACS Omega 5,
plastic waste composition on product yield. J. Anal. Appl. Pyrolysis 51, 39–55. 26005–26014.
Rajasekhar Reddy, B., Vinu, R., 2018. Microwave-assisted co-pyrolysis of high ash Indian Wang, X., Jin, Q., Wang, L., Bai, S., Mikulčić, H., Vujanović, M., Tan, H., 2019.
coal and rice husk: product characterization and evidence of interactions. Fuel Synergistic effect of biomass and polyurethane waste co-pyrolysis on soot formation
Process. Technol. 178, 41–52. at high temperatures. J. Environ. Manag. 239, 306–315.
Rasam, S., Azizi, K., Moraveji, M.K., Akbari, A., Soria-Verdugo, A., 2022. Insights into the Wei, H., Luo, K., Xing, J., Fan, J., 2022. Predicting co-pyrolysis of coal and biomass using
co–pyrolysis of olive stone, waste polyvinyl chloride and Spirulina microalgae blends machine learning approaches. Fuel 310, 122248.
through thermogravimetric analysis. Algal Res. 62, 102635. Wu, B., Guo, X., Qian, X., Liu, B., 2022. Insight into the influence of calcium on the co-
Saeed, S., Saleem, M., Durrani, A., 2020a. Thermal performance analysis and synergistic pyrolysis of coal and polystyrene. Fuel 329, 125471.
effect on co-pyrolysis of coal and sugarcane bagasse blends pretreated by Wu, Y., Zhu, J., Zhao, S., Wang, D., Jin, L., Hu, H., 2021. Co-pyrolysis behaviors of low-
trihexyltetradecylphosphonium chloride. Fuel 278, 118240. rank coal and polystyrene with in-situ pyrolysis time-of-flight mass spectrometry.
Saeed, S., Shafeeq, A., Raza, W., Ijaz, A., Saeed, S., 2020b. Effect of regenerated ionic Fuel 286, 119461.
liquid pretreatment on the thermogravimetric analysis of spent coffee ground. Xu, M., Guan, J., Wang, J.W., 2012. Pyrolysis behavior and kinetics of polyurethane
Energy Sources, Part A: Recovery, Util., Environ. Eff. 1–13. insulation materials from waste refrigerators, advanced materials research, 356-360,
Saeed, S., Saleem, M., Durrani, A., Haider, J., Riaz, M., Saeed, S., Qyyum, M.A., 1752–1758.
Nizami, A.-S., Rehan, M., Lee, M., 2021a. Determination of kinetic and Yao, Z., Yu, S., Su, W., Wu, D., Liu, J., Wu, W., Tang, J., 2020a. Probing the combustion
thermodynamic parameters of pyrolysis of coal and sugarcane bagasse blends and pyrolysis behaviors of polyurethane foam from waste refrigerators. J. Therm.
pretreated by ionic liquid: a step towards optimization of energy systems. Energies Anal. Calorim. 141, 1137–1148.
14, 2544. Yao, Z., Yu, S., Su, W., Wu, W., Tang, J., Qi, W., 2020b. Comparative study on the
Saeed, S., Saleem, M., Durrani, A., Haider, J., Riaz, M., Saeed, S., Qyyum, M.A., pyrolysis kinetics of polyurethane foam from waste refrigerators. Waste Manag. Res.
Nizami, A.-S., Rehan, M., Lee, M., 2021b. Determination of kinetic and 38, 271–278.
thermodynamic parameters of pyrolysis of coal and sugarcane bagasse blends Yao, Z., Cai, D., Chen, X., Sun, Y., Jin, M., Qi, W., Ding, J., 2022. Thermal behavior and
pretreated by ionic liquid: a step towards optimization of energy systems. Energies. kinetic study on the co-pyrolysis of biomass with polymer waste. Biomass-.-.
Saha, G.R., Das, T., Handique, P., Kalita, D., Saikia, B.K., 2018. Copyrolysis of low-grade Convers. Biorefinery.
Indian coal and waste plastics: future prospects of waste plastic as a source of fuel. Yorulmaz, S.Y., Atimtay, A.T., 2009. Investigation of combustion kinetics of treated and
Energy Fuels 32, 2421–2431. untreated waste wood samples with thermogravimetric analysis. Fuel Process.
Seah, C.C., Tan, C.H., Arifin, N.A., Hafriz, R.S.R.M., Salmiaton, A., Nomanbhay, S., Technol. 90, 939–946.
Shamsuddin, A.H., 2023. Co-pyrolysis of biomass and plastic: circularity of wastes Zhang, C., Li, S., Ouyang, S., Tsang, C.-W., Xiong, D., Yang, K., Zhou, Y., Xiao, Y., 2021b.
and comprehensive review of synergistic mechanism. Results Eng. 17, 100989. Co-pyrolysis characteristics of camellia oleifera shell and coal in a TGA and a fixed-
Sharma, S., Ghoshal, A.K., 2010. Study of kinetics of co-pyrolysis of coal and waste LDPE bed reactor. J. Anal. Appl. Pyrolysis 155, 105035.
blends under argon atmosphere. Fuel 89, 3943–3951. Zhang, J., Yang, H., Kang, G., Yu, J., Gao, S., Liu, Z., Li, C., Zeng, X., Lu, S., 2022b. The
Simón, D., Borreguero, A.M., de Lucas, A., Rodríguez, J.F., 2018. Recycling of synergistic effect on the product distribution for the co-pyrolysis of tannery wastes.
polyurethanes from laboratory to industry, a journey towards the sustainability. Fuel 322, 124080.
Waste Manag. 76, 147–171. Zhang, T., Bai, Z., Hou, R., Feng, Z., Ye, D., Guo, Z., Kong, L., Bai, J., Li, W., 2021a.
Singh, R.S., Singh, T., Hassan, M., Larroche, C., 2022. Biofuels from inulin-rich Research progress on co-pyrolysis characteristics of coal and waste plastics. Chem.
feedstocks: a comprehensive review. Bioresour. Technol. 346, 126606. Ind. Eng. Prog. 40, 2461–2470.
Stančin, H., Růžičková, J., Mikulčić, H., Raclavská, H., Kucbel, M., Wang, X., Duić, N., Zhang, T., Yuchi, W., Bai, Z., Hou, R., Feng, Z., Guo, Z., Kong, L., Bai, J., Meyer, B.,
2019. Experimental analysis of waste polyurethane from household appliances and Li, W., 2022a. Insight into the charging methods effects during clean recycling of
its utilization possibilities. J. Environ. Manag. 243, 105–115. plastic by co-pyrolysis with low-rank coal. J. Clean. Prod. 333, 130168.
Stančin, H., Mikulčić, H., Manić, N., Stojiljiković, D., Vujanović, M., Wang, X., Duić, N., Zhang, T., Bai, Z., Guo, Z., Kong, L., Dou, H., Bai, J., Li, W., 2023. Synergistic effect of
2021. Thermogravimetric and kinetic analysis of biomass and polyurethane foam Hami coal and polystyrene on products properties during co-pyrolysis: Evaluation on
mixtures Co-Pyrolysis. Energy 237, 121592. the contribution of different interaction sources. Fuel 338, 127209.
Stančin, H., Šafář, M., Růžičková, J., Mikulčić, H., Raclavská, H., Wang, X., Duić, N., Zhou, L., Zhang, G., Zhang, L., Klinger, D., Meyer, B., 2019. Effects of contact conditions
2022. Influence of plastic content on synergistic effect and bio-oil quality from the between particles and volatiles during co-pyrolysis of brown coal and wheat straw in
a thermogravimetric analyzer and fixed-bed reactor. Processes 7, 179.

921

You might also like