Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

Journal Pre-proof

Kinetic and thermodynamic analysis of Putranjiva roxburghii (Putranjiva) and Cassia


fistula (Amaltas) non-edible oilseeds using thermogravimetric analyzer

Abhisek Sahoo, Sachin Kumar, Kaustubha Mohanty

PII: S0960-1481(20)31742-0
DOI: https://doi.org/10.1016/j.renene.2020.11.011
Reference: RENE 14451

To appear in: Renewable Energy

Received Date: 8 June 2020


Revised Date: 4 October 2020
Accepted Date: 4 November 2020

Please cite this article as: Sahoo A, Kumar S, Mohanty K, Kinetic and thermodynamic analysis
of Putranjiva roxburghii (Putranjiva) and Cassia fistula (Amaltas) non-edible oilseeds using
thermogravimetric analyzer, Renewable Energy, https://doi.org/10.1016/j.renene.2020.11.011.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Credit Author Statement

Abhisek Sahoo: Conceptualization, Formal analysis, Investigation, Methodology, Software,

Visualization, Data Curation, Writing – Original Draft. Sachin Kumar: Project

administration, Supervision, Validation, Resources, Writing – Review & Editing. Kaustubha

Mohanty: Data Curation, Supervision, Writing – Review & Editing.

of
ro
-p
re
lP
na
ur
Jo
Jo
ur
na
lP
re
-p
ro
of
1 Kinetic and thermodynamic analysis of Putranjiva roxburghii (Putranjiva) and Cassia
2 fistula (Amaltas) non-edible oilseeds using thermogravimetric analyzer

3 Abhisek Sahoo1, Sachin Kumar 1, 2,*, Kaustubha Mohanty3


1
4 Department of Energy Engineering, Central University of Jharkhand, Ranchi, India – 835205

2*
5 Centre of Excellence – Green & Efficient Energy Technology (CoE-GEET), Central University of Jharkhand,
6 Ranchi, India - 835205

3
7 Department of Chemical Engineering, Indian Institute of Technology Guwahati, Guwahati – 781039, India

of
ro
10

11 -p
12
re
13
lP

14
na

15
ur

16
Jo

17

18

19

20

1
21 – sahooabhisek01@yahoo.com

3
22 – kmohanty@iitg.ac.in

*
23 Author to whom correspondence shold be addressed

24 Electronic mail: sachin.kumar.01@cuj.ac.in

25

26

1
27 Abstract:

28 Kinetic triplets and thermodynamic parameters of the pyrolysis of Putranjiva roxburghii (PR)
29 and Cassia fistula (CF) non-edible oilseeds were studied using thermogravimetric analyzer.
30 The kinetic parameters were assessed using both model-free [Kissinger-Akahira-Sunose
31 (KAS), Flynn-Wall-Ozawa (OFW), Starink (STR), Li and Tang (LTA), Friedman (FRM),
32 Vyazovkin (VYZ), Kissinger (KN), Avrami (AVM), and Master-plot (MP)] and model-
33 fitting [Coats-Redfern (CR)] methods at four different heating rates (10, 25, 40, and 55
o
34 C/min) somewhere in the range of 10% and 80% conversions. The reaction mechanisms
35 were found to be in good agreement with the experimental thermal analysis data.
36 Thermodynamic parameters (∆G, ∆H, and ∆S) are also determined by model-free

of
37 isoconversional method. The kinetic and thermodynamic parameters recommended the

ro
38 appropriateness of PR and CF non-edible oilseeds for pyrolysis process.

39 -p
re
40 Keywords: Putranjiva and amaltas seed; Non-isothermal pyrolysis kinetics; TGA analysis;
Reaction mechanism; Thermodynamic analysis.
lP

41

42
na

43 1. Introduction
ur

44 Biomass is one of the most recognized sustainable resources for the production of renewable
Jo

45 and clean energy due to its capability to control the amount of toxic gases and carbon
46 emission in the environment. Biomass is a biological material which mainly contains carbon,
47 oxygen, hydrogen, least amount of sulphur and ash. Among all kinds of biomass, some
48 contain a notable amount of inorganic species. Due to geographical locations and climate
49 conditions, some variation has occurred in biomass composition and their properties [1].
50 India has a rich diversity of about 150 species of plant yielding, which can be the production
51 of fuels and other sources of energy. In the world, the third-largest source of energy is
52 biomass. The biomass availability of different sources of the globe is about 220 billion tons
53 per year [2]. Lignocellulosic biomass is a plenteous and inexpensive source of renewable
54 energy. Putranjiva roxburghii and Cassia fistula non-edible oilseeds are the examples of
55 lignocellulosic biomass available in India. Putranjiva roxburghii (Putranjiva) is an evergreen
56 tree of height up to 12m. Seeds of this plant usually stone pointed and very hard. The
57 extracted oil from this seed is used for renewable fuel while blending up to 30-40% with

2
58 diesel. It is distributed in various regions, including in India, Thailand, Nepal, Bangladesh,
59 Myanmar, and Sri Lanka. Cassia fistula (Amaltas) is a deciduous and medium-sized tree up
60 to 24 m in height and 1.8 m in width. Seed diameter of this plant is about 0.25 to 0.5 inches
61 with a flattened, smooth, and teardrop in shape. These are mildly poisonous seeds, and it can
62 cause diarrhoea and vomiting when ingested. The tree is one of the most wide-spread in the
63 Indian forest particularly in South India and Central India. It is also widely spread in Asia,
64 China, Brazil, West Indies, and South Africa [3]. These non-edible oilseeds have not found
65 very useful applications. Pyrolysis may be used as a potential thermochemical conversion
66 process for effective handling of these oilseeds. Based on this motivation, the present work
67 emphasized on the pyrolysis kinetics of these oilseeds.

of
68 The kinetic studies of any feedstock under pyrolysis conditions give essential

ro
69 information on how it disintegrates. It also reveals the fundamental information on the
70
71
-p
reaction mechanisms and build up a numerical model which could depict the process [4].
Understanding the estimation of kinetic parameters furthermore assists with computing the
re
72 expenses of the thermal process, such as pyrolysis [5]. Kinetic investigations on the biomass
lP

73 pyrolysis are extremely repetitive in the literature, and tended by using the two
74 methodologies: i) model-fitting methods, and ii) model-free methods [6, 7]. Isoconversional
na

75 methods can assess the behavior of the complex reaction; but it’s limitation is the threat of
76 selecting an unsatisfactory kinetic model and finding an inappropriate kinetic parameter [8].
ur

77 The pyrolysis kinetics and possible reaction mechanism of lignocellolosic biomass by the
Jo

78 thermogravimetric analyzer are of great importance to study. Knowledge of kinetics derived


79 from such experiments is very significant for the optimal design and operational settings of a
80 thermochemical conversion reactor [9]. Pyrolysis kinetic behaviour of Samanea saman seeds
81 (SS) was carried out using Kissinger, Distributed Activation Energy Model (DAEM) and
82 Miura-Maki-Integral (MMI) method. Kinetic results confirmed that the average activation
83 energy was found 118.24 kJ mol-1, 168.70 kJ mol-1, and 97.87 kJ mol-1 for Kissinger, DAEM,
84 and MMI model respectively [10]. Thermogravimetric analysis of Acai (Euterpe oleracea
85 Mart.) seed biomass was performed to determine the kinetic and thermodynamic parameters
86 of the pyrolysis process for its energy recovery. Kinetic parameters were determined using
87 Flynn-Wall-Ozawa (FWO) and Kissinger-Akahira-Sunose (KAS) iso-conversional models.
88 The activation energy of acai seed was estimated 159.12 kJ/mol and 157.62 kJ/mol by FWO
89 and KAS, respectively. Thermodynamic parameters, such as ∆H, ∆G and ∆S, and the high
90 heating value (21.1 MJ kg−1) showed that the acai seed is an excellent potential source of

3
91 raw material for bioenergy production [11]. Kinetic investigation of Putranjiva roxburghii
92 (PR), Cassia fistula (CF) and Albizia lebbeck (AL) non-edible oilseed samples was carried
93 out by using single heating rate models such as Broido and Horowitz-Metzger methods. The
94 activation energy (Ea) observed was in the order of PR > AL > CF by both the methods. The
95 maximum activation energy was found in PR (34.46 KJ/mol by BR method and 56.26 KJ/mol
96 by HM method), followed by AL and CF [3].

97 The literature review on non-isothermal pyrolysis kinetics of PR and CF has


98 revealed that there is no research work on account of possible pyrolysis reaction mechanism
99 and apparent activation energy by advanced model-free and model-fitting methods. The
100 possible reaction path and actual kinetic parameters have many advantages to optimize the

of
101 appropriate reactor design for the industrial and scale-up applications. To the best of our

ro
102 knowledge, it is the first study to report possible reaction mechanisms, apparent activation
103
104
-p
energy and thermodynamic parameters by some advanced kinetic models such as model-free
and model-fitting methods. There is no evidence in the literature on using the different non-
re
105 isothermal approaches to determine the kinetic parameters for Putranjiva roxburghii and
lP

106 Cassia fistula under the inert atmosphere.

107 The primary objective of this work is to analyse the thermal decomposition kinetics for
na

108 describing the process of devolatilization and to estimate the thermodynamic parameters. In
109 the present work, the non-isothermal pyrolysis of PR and CF was performed in a
ur

110 thermogravimetric analyzer, through which the weight loss to the temperature curve
Jo

111 (degradation curve) could be obtained. The degradation curves were further analyzed by
112 model-free KAS, OFW, STR, LTA, FRM, VYZ, KN, AVM, MP methods and model-fitting
113 CR method for kinetic analysis involving changes in the apparent activation energy (Ea). The
114 thermodynamic parameters (∆H, ∆G, and ∆S) were assessed by model free isoconversional
115 method.

116 2. Materials and methods


117 2.1. Feedstocks and their characterization

118 Two non-edible oilseeds – Putranjiva roxburghii (Putranjiva) and Cassia fistula (Amaltas)
119 were used in this work. After collection, the oilseeds were first undergone sun-dried for 24h
120 to remove excessive amount of moisture and subsequently, crushed into powdered form (size
121 > 2mm). The proximate, elemental, and heating values of these seed samples were
122 determined using standard test procedures. Raw non-edible oilseeds (without extraction) are

4
123 used for all the analysis except the compositional analysis. The detailed characterization of
124 these oilseeds are found in our earlier publication [3].

125 2.2. Thermogravimetric analysis

126 TGA (NETZSCH) was carried out under N2 gas atmosphere (flow rate: 40ml/min for purge
127 and 20ml/min for protective respectively) at four different heating rates (10, 25, 40, and 55
o
128 C/min) for a temperature range of 25 – 925 oC. About 7.5 ± 0.03 mg of seed sample was
129 placed on an aluminum crucible and held to equilibrate at 25 oC. Each experiment was
130 revised at least twice to ensure the reproducibility of the data. The DTG data for all the seed
131 samples were taken from the TGA data utilizing Origin Pro software®.

of
132 2.3. Data analysis of TGA

ro
133 The parameters of thermal-decomposition, statistical parameters (R2), kinetic modeling, and
134
135
-p
their analysis were carried out using Microsoft Office Excel 2016 (version
16.0.12325.20344), RStudio (version 1.2.5033), and Origin 2020 (Learning Edition)
re
136 softwares.
lP

137 The experimental TGA data and its derivative data (DTG) became first standardized by
138 utilizing eqs. (1) and (2) individually to normalized the four heating rate programs. The
na

139 standardized TGA and DTG data are represented by M and dM/dt, respectively.
ur

140 = (1)
Jo

141 = = (2)

142 Whereas mo, and mt is the initial mass at 20.0 ± 2.5 oC, and the particular mass at a time (t),
143 respectively, wherein dmt/dt corresponds to the experimental DTG data.

144 However, the conversion of solid-gas reaction is usually attained


145 experimentally as a function of physical assets, for instance, the weight loss, and released or
146 absorbed of the heat. Therefore, the values of conversion and conversion rate usually
147 determined via following eqs. (3) and (4), respectively. Conversion and conversion rate is
148 represented by α and dα/dt, respectively, where Mo, Mt, and M∞ are the initial mass, mass at a
149 given time (t), and final mass, respectively, for each step at various heating rate programs.
150 The kinetic parameters (activation energy, reaction mechanism, pre-exponential factor, and
151 reaction order) have been calculated, considering each conversion step aside.

5
152 = (3)

153 =− (4)

154 2.4.Pyrolysis performance index (PPI)

155 Based on the TG-DTG thermographs, the pyrolysis performance index estimated in the
156 present study included ignition temperature (Ti) and time (ti), peak temperature (Tp) and time
157 (tp) corresponding to the maximum degradation rate (-Rp), burnout temperature (Tb) and time
158 (tb), and mean degradation rate (Rv). The five pyrolysis performance indices such as;
159 devolatilization index (Dv), ignition index (Di), burnout index (Db), combined combustion

of
160 index (S), and flammability index (C) were determined as follows [12, 13]:

ro
= (5)
× ×∆ .
161

=
-p (6)
×
162
re
=∆
lP

163 ! (7)
. × × "

×
#=
na

164 $× (8)
"

%=
ur

165 $ (9)
Jo

166 Whereas ∆T0.5 and ∆t0.5 is the temperature and time interval at the half value of the Rp,
167 respectively.

168 2.5. Kinetic theory

169 The composition (Hemicellulose, Cellulose, and Lignin) of biomass differ in diverse
170 lignocellulosic biomass types. Therefore, biomass pyrolysis is a complex process due to the
171 inclusion of a few thermochemical reactions. Consequently, predicting an exact pyrolysis
172 kinetic reaction is somewhat tricky. However, a wide-ranging global pyrolysis reaction
173 mechanism of biomass can be inscribed as:

.
174 &'()*++ +(,'- /01 2(,*3',4+ 5(6-46+*7,4 + 6(65(6-46+*7,4 +
175 %ℎ*: +(,'- :4+'-;4 (10)

176 A solid-gas conversion (heterogeneous reaction) rate can be expressed as:

6
177 =< = (11)

178 Whereas k(T) is the reaction rate constant, which is based on the temperature program, and
179 f(α) is the reaction mechanism.

180 From the eq. (11) the rate of conversion (dα/dt) depends on the reaction rate constant k(T)
181 and the reaction mechanism f(α). Thus, the reaction rate constant k(T) can be expressed by an
182 Arrhenius eq as:
@AB
183 < = >4 ? CD E (12)

184 Whereas Eα is the apparent activation energy (kJ/mol), T is the absolute temperature (K), A is

of
185 the pre-exponential factor (min-1), and R is the universal gas constant (8.314 J/K.mol).

ro
186 By combining eqs. (11) and (12) we get:

187 = >=
@AB
4 ? CD E
-p (13)
re
188 In the pyrolysis process, the heating rate (β, oC/min) plays a vital role in estimating the
lP

189 kinetic parameters. Thus, it can be expressed in terms of conversion rate (dα/dt) as:

F= =
na

190 (14)
ur

191 Utilizing eq. (14) in eq. (13) then it becomes:


@AB
G ? E
= H= 4
Jo

192 CD (15)

193 The kinetic analysis of biomass mostly determined by isoconversional


194 methods, which classified by the differential and integral methods with different temperature
195 integral approximations. The differential method can be directly calculated from the eq. (15),
196 it is called Friedman method. However, the integral methods are performed from reordering
197 and integration of both sides of eq.(15), along with eq. (16):

@AB
G GLB
I =J O
= HJ O
4 ? CD E - = M N (16)
K H
198

199 The left-hand side g(α) indicates the integral form of the reaction models while the right-hand
200 side signifies the temperature integral without an exact analytical solution. However,
201 typically, it can be resolved by empirical approximations, p(x), where x=E/RT.

202

7
203 2.6. Kinetic parameters by Model-free methods

204 All isoconversional methods depend upon the principle that at a constant value of the degree
205 of conversion (α), the rate of reaction along with temperature (T) only [14].

206 2.6.1. The activation energy (Ea) by Isoconversional methods

207 In this present study, the six isoconversional methods such as Kissinger-Akahira-Sunose
208 (KAS), Flynn-Wall-Ozawa (OFW), Starink (STR), Li and Tang (LTA), Friedman (FRM) and
209 Vyazovkin (VYZ) represents by eqs. (17) – (27), respectively, were utilized. The subscript ‘i’
210 in the following eqs. indicates the different heating rate programs, and Tα,i is the temperature
211 related to each degree of conversion (α).

of
212 2.6.1.1. Kissinger-Akahira-Sunose (KAS) method

ro
213 The KAS method is based on the following temperature integral approximation p(x) given by
214 Murray and White [15]:
-p
re
Q @R
M N ≅ (17)
S$
215
lP

216 Substitute the value of eq. (17) on eq. (16), then we get:
na

H GLB LB
ln V $ X = Y,6 [− (18)
Z B,
217
B,
ur

H
The eq. (18) is recognized as the KAS method. Eα can be calculated by the slope of ln ? $ E
B,
218
Jo

versus for each value of conversion (α).


B,
219

220 2.6.1.2. Flynn-Wall-Ozawa (OFW) method

221 The OFW method is based on Doyle’s approximation of temperature integral [16]:

222 M N ≅4 \.] ^ _.`^abS


(19)

223 Substitute the value of eq. (19) on eq. (16), then it becomes:

GLB LB
ln F = ln Y Z
[ − 2.315 − 0.4567 (20)
B,
224

225 The eq. (20) is known as the OFW method. Eα can be evaluated by the slope of ln F versus

for each value of conversion (α). However, the OFW method is less accurate than the
B,
226

227 KAS method [6].

8
228 2.6.1.3. Starink (STR) method

229 The STR method is based upon the following temperature integral approximation [17]:

Q @k. lR _.] \
M N ≅ (21)
S k.m$
230

231 By putting the above approximation value on eq. (16), then we get:

H LB
232 ln V k.m$ X = Constant − 1.0008 V X (22)
B, B,

233 The above eq. (22) is known as the STR method. Activation energy obtained by this method
234 is more accurate than OFW and KAS methods. The activation energy (Ea) can be determined

of
H
by the slope between ln V k.m$ X versus for each value of conversion (α).
B,
235
B,

ro
236 2.6.1.4. Li and Tang (LTA) method

237 The LTA method is based on the following equation without any temperature integral
-p
re
238 approximation [18]:
lP

LB
239 J_ ln ? E = − J_ ? E - (23)
, ,
na

240 Where = ,6> + J_ ln[= ]-


ur

241 The plot between J_ ln ? E versus J_ ? E - gives the slope to determine the
, ,
Jo

242 activation energy (Eα).

243 2.6.1.5. Friedman (FRM) method

244 The FRM method is based on differential methods without any approximation eq. This
245 method has diminished the probability of error. The following eq is known as FRM method
246 [19]:

Lx
247 ln ? E = ln vF ? E w = ln[> = ]− (24)
, , , B,

The slope lines between ln ? E versus at each conversion (α) are used for determining
B,
248
,

249 the activation energy (Eα).

250

9
251 2.6.1.6. Vyazovkin (VYZ) method

252 The VYZ method is a non-linear approach. The activation energy (Eα) was evaluated by
253 minimizing the following eq. [20]:

|}Lx , B, ~H•
y z = ∑€ƒ ∑€•‚ =6 6−1 (25)
|}Lx , B,• ~H
254

255 here, n is the number of experiments at different heating rates.

@AB
? E L
256 Where „ z… , = J_ B 4 CD - = ? BE M N (26)

257 In the above equations „}z… , , ~ and „}z… , ,• ~ represents the integral temperature M N

of
258 corresponding to the heating rate programs F and F• , respectively. „ z… , has no

ro
259 analytical solution, it can be evaluated by direct numerical integration or by utilizing the
260 Senum-Yang approximation. In this work, the following the 8th order rational approximation
261 was used [21]:
-p
re
Q @R S † ‡b_S ˆ ‡ ‰‰aS ‡\`Š\_S ‹ ‡ b_ ]aSŒ ‡^bb^‰`S $ ‡‰``^a_S‡]^b \_
262 M N = S$
?S l ‡b\S† ‡\_\`Sˆ ‡\‰^a_S ‡\ ab\_S ‹ ‡‰‰_]\_SŒ ‡ bŠ`\`_S $ ‡ ^b\`‰_S‡`_]\__
E (27)
lP

LB
263 Where, N =
na

264 2.6.2. Reaction mechanism [g(α)] by Master plot (MP) method


ur

265 The master plot method is one of the most direct and straightforward alternatives to evaluate
Jo

266 the reaction model of the solid-state conversion process. According to the ICTAC, the
267 theoretical reaction models [Table – 1] were evaluated via the master plot method [22]. The
268 experimental master plots were based on the normalized eq. (28). In the present work, the
269 following eq. was utilized to evaluate the reaction mechanism:

AB •
Z • S B • S B
= =•
ŽC
AB • (28)
Z S .
270
. • S .
ŽC

Z • S B
271 Where Z and • S .
are the theoretical and experimental master plots, respectively.
.

272

273

274

10
275 Table 1. 22 kinetic models (f(α)) and their integral expression (g(α)) of the solid-state
276 pyrolysis process

277

= -
I = •
Symbol
_ =
Geometrical Contraction Model
R1 1
2 1− 1− 1−
R2 \ \
\
3 1− 1− 1−
R3 ] ]
Diffusional Model
D1 1 \

of
D2 [− ln 1 − ] 1− ln 1 − +
3 \

ro
\
1− ] v1 − 1− ]w v1 − 1 − ]w
D3
2
3 -p 2 \
v 1− − 1w 1−V X − 1−
D4
] ]
2 3
re
Nucleation Model
3 \
1− [− ln 1 − ]] [− ln 1 − ]]
A1.5
2
lP

2 1− [− ln 1 − ]\ [− ln 1 − ]\
A2
\
3 1− [− ln 1 − ]] [− ln 1 − ]]
A3
na

]
4 1− [− ln 1 − ]` [− ln 1 − ]`
A4
4 ^
[− ln 1 − ]`
ur

1− [− ln 1 − ]
A4/5
`
5
5 a
1− [− ln 1 − ] [− ln 1 − ]^
Jo

A5/6
^
6
Order Based Model
F1 1− [− ln 1 − ]
]
1− 2v 1− − 1w
F1.5 \ \

F2 1− \ [ 1− − 1]
F3 1− ] [ 1− \
− 1]
2
F4 1− ` [ 1− ]
− 1]
3
Power Law Model
P1.5 3 \
] ]
2
2
P2 \ \
\
3
P3 ] ]
]
4
P4 ` `

11
278 2.6.3. Pre-exponential factor (A) by Kissinger (KN) Method

279 The KN method is employed for evaluating the pre-exponential factor by utilizing the above
280 kinetic parameters. The proposed equation of the Kissinger method is [23]:

H L G
ln V $ X = − + ln ? E (29)
L
281

282 On arranging the above equation for calculating the pre-exponential factor (A), then the
283 equation will become:

A
‘ x ’
CD
HLx Q
284 >= $ (30)

of
285 Where, • is the peak temperature of the DTG curve and z… is the activation energy, which is

ro
286 obtained by the VYZ method.

287 2.6.4. Order of reaction (n) by Avrami (AVM) theory


-p
re
288 The Avrami theory can be utilized to evaluate the order of reaction (n) from a linear
relationship between the ln − ln 1 − and ln F described as [24]:
lP

289

LB
ln ln 1 − = ln > − − 6,6 F
na

290 (31)

291 For a particular temperature (T), the points of ln − ln 1 − versus ln F at different


ur

292 heating rates (β) can be fitted to a straight line, and the slope of the line corresponds to -n.
Jo

293 Therefore, the order of reaction (n) can be calculated from the slope of the straight line.

294 2.7. Kinetic parameters by Model-fitting methods

295 In the model-fitting isoconversional analysis, there are various model-fitting methods have
296 been used in a wide range of studies. However, in this work, the Coats-Redfern (CR) method
297 was employed to demonstrate the kinetic triplets [Ea, A, and g(α)], and it also evaluates the
298 overall order of reaction (n). While choosing the order of reaction (n), the reaction rate can be
299 expressed as [25]:

AB
G V X
300 = H4 CDB,
1− €
(32)

301 On arranging the above, then the equation becomes,

12
AB
G V X

= 4 CDB,
- (33)
H
302

303 Integrate the above then it becomes,

AB
k@“ G V X
= J_ B, 4 CDB,
- (34)
€ H
304

305 The above integral part has no analytical solution; by employing the asymptotic series and
306 neglecting the higher-order terms, now it will become: -
k@“ G \ L
ln ” $
• = ln ” ?1 − E• − [for n ≠ 1] (35)
€ HL L
307

–— G \ L

of
308 ln ” $ • = ln ”HL ?1 − L
E• − [for n = 1] (36)

ro
\
309 However, the term L
<< 1; hereupon it can be omitted. Considering this assumption, the

310 above eqs. (35) and (36) can be modified, so the above explanation will become:-
-p
re
k@“ G L
311 ln ” $ €
• = ln ”HL• − [for n ≠ 1] (37)
lP

–— G L
312 ln ” $
• = ln ”HL• − [for n = 1] (38)
na

313 Whereas T = absolute temperature (oC), β = heating rate program (oC/min), R = gas constant
ur

314 (8.314 J/mol.K), E = activation energy (kJ/mol) and A = pre-exponential factor (min-1), and n
315 = order of reaction.
Jo

316 When computing the reaction mechanism g(α), by combining the eq. (32) and CR equation
317 [26], the g(α) can be expressed: -

G $ L
I = exp ?− E (39)
LH
318

319 On arranging the above, the equation becomes: -

–—[Z ] G Lx
320 $ = ln YHL[ − (40)
B B

Z
321 Plotting the left-hand side of eq. (40), which includes ln[ $ ] vs. , gives Ea and A from the
B B

322 slope and intercept, respectively. The g(α) can be chosen from table – 1. The model that gives
323 the best coefficient of determination (R2) with similar Ea value, which was obtained by the
324 average value of the model-free method (VYZ), is selected as the appropriate model.

13
325 2.8.Thermodynamic parameters

326 From the kinetic parameters via above model-free methods and degradation data by TGA are
327 likely to obtain the following thermodynamic parameters of pyrolysis process such as; Gibbs
328 free energy (∆G), enthalpy (∆H), and entropy (∆S). These parameters can be evaluated
329 depending on the following equations [13]:

.œ •
∆ = z… + › ln ? E (41)
žG
330

331 ∆Ÿ = z… − › (42)

∆ ∆¡
332 ∆# = (43)

of
333 Whereas kB is the Boltzmann constant (1.381×10-23 J/K), h is the Plank constant (6.626×10-34

ro
334 J/s), A is the pre-exponential factor (s-) which was calculated from KN method, Tm is the
335 -p
peak temperature at the corresponding heating rate, and Eα is the activation energy (kJ/mol) at
a particular conversion (α) by VYZ method.
re
336

337 3. Results and discussion


lP

338 3.1. Feedstocks characterization


na

339 Table – 2 shows the results of the physicochemical characterization of the Putranjiva
340 roxburghii (PR) and Cassia fistula (CF) non-edible oilseeds and its comparison with other
ur

341 oilseeds available in the literature. The carbon and oxygen fractions of PR and CF indicated
Jo

342 that these non-edible seeds had a higher organic compound and calorific value. The sulfur
343 content of PR and CF (far below 1%) pointed towards the lesser SOx emissions. From the
344 proximate analysis, the volatile content were found to be 86.25% and 79.65% which were
345 followed by the ash content (5.68%) and (6.03%), while the moisture contents were (3.81%)
346 and (6.26%) in PR and CF oilseeds respectively. This further suggested that PR and CF
347 oilseed samples were easy to decompose. From Table 2, it also indicated that the nitrogen
348 percentage and fixed carbon content in PR sample were much lower than CF. Higher the
349 calorific value 25.14 MJ/kg and 18.68 MJ/kg of PR and CF, respectively, than the original
350 coal (10.56 MJ/kg) makes these oilseeds a favorable prospect for cleaner energy production
351 [27].

352

14
353 Table 2. Results of the physicochemical characterization of PR and CF, along with other
354 reported biomass materials on the air-dried basis (adb).

Sample Proximate analysis (wt%) Elemental analysis (wt%) Extractives HHV Reference
(wt%) (MJ/kg)
MC AC VM FC* C H N S O*
PR 3.81 5.68 86.25 4.26 51.96 9.24 0.74 0.86 37.19 50.55 25.14 Present
study
CF 6.26 6.03 79.65 8.06 42.37 7.51 2.98 0.39 46.74 18.83 18.68 Present
study
SS 6.19 3.06 76 14.74 48.46 6.75 7.3 - 37.47 30.66 17.68 [28]
LS 6.53 10.5 76.95 6.02 61 8.5 3 0.3 27.2 33.55 27.9 [29]
AM 2.9 4.88 87.88 4.34 53.33 7.541 3.255 0.90 34.9 - 9.55 [30]
NS 16.0 4.95 71.0 8.05 38.42 8.27 7.48 0.69 45.14 - 23.01 [31]
*
355 Calculated by difference; FC = 100 – (MC+AC+VM), and O = 100 – (C + H + N + S)

356 3.2. TG-DTG analysis

of
357 The degradation curves of two non-edible oilseed have been obtained using

ro
358 thermogravimetric analyzer (TGA). Fig. 1(a) and (b) represents the weight loss (%) along
359
360
-p
with temperature (oC) during the thermal decomposition of PR and CF oilseeds at different
heating rates under N2 atmosphere. In general, the thermal degradation curves of
re
361 lignocellulosic materials can be partitioned into three zones, which are impacted by various
lP

362 physiochemical composition in terms of hemicellulose, cellulose, and lignin. The above three
363 zones, such as i) dehydration ii) devolatilization and iii) char formation zone. Similarly, these
na

364 three degradation zones were likewise seen in other oilseeds also such as; samanea saman
365 seeds [28] and pongamia pinnata seeds [29] with particulates in the temperature range.
ur
Jo

15
of
ro
-p
re
366
lP

367 Fig. 1(a). Thermogravimetric (TG) curves of PR


na
ur
Jo

368

369 Fig. 1(b). Thermogravimetric (TG) curves of CF

16
370 TG thermograph [Fig. 1] confirms that the first zone or dehydration zone started from
371 room temperature (25 ± 2.5 oC) and ended at about 170 ± 10 oC, the mass loss in this zone is
372 due to the evaporation of moisture content and hydrolysis of low molecular weight
373 compounds such as extractives [32]. Besides, the normalized weight loss variations were not
374 significant in between 170 – 200 oC. The second zone or devolatilization zone which started
375 from 200 oC to 450 oC for both the oilseeds. This zone is the main decomposition region,
376 where volatile materials were released due to the breakdown of hemicellulose, cellulose, and
377 fractional decomposition of lignin. Depending upon the biomass composition, the
378 temperature range in this region may slightly vary with weight loss of biomass. The third
379 zone or char formation zone, occurred at a higher temperature (> 500 oC) at a slower rate of

of
380 decomposition because of lignin linked with the group of phenolic hydroxyl [33]. At this

ro
381 stage, the leftover mass, mostly made out of carbonaceous solid formed in an inert
382 atmosphere, was pyrolyzed and discharged gases. Moreover, as seen in [Fig. 1(a) and (b)],
383
-p
the trend of the degradation curve of PR and CF are not very much influenced by heating rate
re
384 programs of 10, 25, 40 and 55 oC/min; there is exclusively a slight shift to the upper zone in
385 the temperature range from 200 – 450 oC, which is also known as active pyrolysis region.
lP

386 DTG thermograph [Fig. 2] confirms that the first peak was observed at the temperature
na

387 range 150 – 200 oC due to the evaporation of moisture and light volatile matters in both the
388 oilseeds. From the [Fig. 2], it also reveals that the DTG thermograph of PR and CF with four
ur

389 heating rates in N2 atmosphere overlapped on each other due to the endothermic process and
instability in thermographs below 200 oC. The second and third peaks in both the oilseeds
Jo

390
391 appeared due to the decomposition of hemicellulose and cellulose compounds. In DTG
392 thermographs of PR [Fig. 2(a)], the decompositions were symbolized by a downslope
393 showing mainly the extractives (waxes, fats, simple sugars, pectins, and phenolic
394 compounds), which differ in types of biomass, even varying among different parts of the
395 respective plant [34]. PR has much higher content of extractives (50.55%) than CF. From
396 [Fig. 2 (b)], the thermal degradation of CF is mainly referred to as the degradation of
397 hemicellulose, cellulose, and lignin. In both the DTG curves [Fig. 2(a) and (b)], a lower
398 temperature shoulder characterizes at various heating rates of both the oilseeds at around 325
o
399 C and 250 oC for PR, and CF, respectively corresponding to the hemicellulose
400 decomposition, and a higher temperature peak around 450 oC and 350 oC for PR, and CF
401 respectively due to the depolymerization reaction and cellulose devolatilization. According to
402 the available literature, it was revealed that the cellulose decomposition occurs in two ways:

17
403 i) depolymerization reaction with the minor formation of CO, CO2 and carbonaceous residues
404 at low temperatures; and ii) integration of bonds at higher temperature besides the significant
405 formation of the bio-oil possessing an extensive range of organic compounds [35]. Finally, at
406 the end of a long linear tail of DTG thermographs (> 500 oC), lignin decomposed at a slower
407 rate and yielded maximum char and released gases. Above the 600 oC, the DTG curve was
408 stagnant, signifying the termination of the pyrolysis reaction of the residual carbonaceous
409 solid. However, as shown in [Fig. 2], DTG thermographs shows an increase in maximum
410 degradation rates and a minor alteration of the significant peak to higher temperatures
411 followed by higher heating rate programs, mainly caused by the mutual impacts of the heat
412 and mass transfer process at various heating rate programs [36].

of
ro
-p
re
lP
na
ur
Jo

413

414 Fig. 2 (a). Differential thermogravimetric (DTG) curves of PR

18
of
ro
-p
re
lP

415

416 Fig. 2 (b). Differential thermogravimetric (DTG) curves of CF


na

417 By comparing TG-DTG thermographs of PR and CF [Fig. 1 and 2], it was observed
ur

418 that the different trends of the TG-DTG thermographs of the non-edible oilseeds are
419 attributed to the various physical and chemical composition; notably hemicellulose, cellulose
Jo

420 and lignin, and some extractive components [3].

421 3.3. Pyrolysis performance index (PPI) analysis

422 Ignition temperature (Ti), peak temperature (Tp), and burnout temperature (Tb) gradually
423 moved to a higher temperature zone when the increase in heating rates under N2 atmosphere,
424 due to thermal hysteresis effect. The effect of heating rate programs on the pyrolysis
425 performance index is presented in [Table – 3]. It was observed that the devolatilization index
426 (Dv) of PR is higher (22.30×10-6 %K-3.min-1) at a higher heating rate (55 oC/min) as
427 compared to CF (10.49×10-6 %K-3.min-1) at the corresponding heating rate. The higher value
428 of Dv implies the better release of the volatiles during the pyrolysis process. The heating rate
429 has a direct relation to Dv. When the heating rate increased, the value of Dv is also increased
430 by around 25 times in both the oilseeds. Similarly, the Di and Db values were significantly
431 higher in the PR than CF. Consequently, the higher heating rate at the N2 atmosphere had all
19
432 the earmarks of being the most useful for the volatilization of both non-edible oilseeds. The
433 value of Dv, Di, Db, S, and C were increased rapidly when the heating rate was increased up
434 to some extent. All the pyrolysis index value of PR was higher than of CF due to high volatile
435 matter and high extractive content present in the PR. Moreover, the ignition temperature (Ti)
436 revealed that PR decomposition started at a lower temperature (180 oC) followed by CF (203
o
437 C). As the value of heating rate increased, the value of ignition (Di) and burnout (Db) indices
438 in both the cases increased very quickly as seen in the Table – 3. Therefore, it was observed
439 that increase in the heating rate prompted the concentrated and intensified pyrolysis process
440 for PR and CF. The values of combined combustion index (S) are increased with increase in
441 the heating rate; it is suggested that the better performances of ignition and burnout estimated

of
442 on account of PR than CF. The flammability index (C) of PR is much higher than the CF due

ro
443 to the lower moisture content and higher the heating value in the PR sample than CF. From
444 table – 3, the difference between the peak temperature (Tp) of PR and CF is around 80 oC.
445
-p
The total weight loss (Wt) of PR and CF are not very much influenced by changing the
re
446 heating rate programs from 10 to 55 oC/min. It is observed that the burnout temperature (Tb)
447 for PR and CF pyrolysis is in the range from 438 oC to 462 oC at different heating rate
lP

448 programs. Similarly, the range of burnout temperature for incense sticks pyrolysis under
449 nitrogen atmosphere was found to be 766.8 oC - 861.8 oC at different heating rate programs
na

450 [10].
ur

451 Table 3. Pyrolysis performance indices of PR and CF with four heating rates program
Jo

Samp Heati Ti Tb Tp ti tf tp Rp Rv Wt Dv Di Db S C
le ng (oC) (oC) (oC) (mi (min (mi (%/mi (%/mi (%) (102) (104 (107 (105
Rate n) ) n) n) n) ) ) )
10 179.4 442.8 401.2 15.4 40.72 37.6 5.57 0.934 84.0 0.88 0.96 0.44 3.65 17.3
PR 4 3 5 5 5 5 7 0
25 187.7 450.9 418.1 6.52 17 15.7 14.55 2.347 84.5 5.83 14.16 7.28 21.4 41.2
9 0 8 6 2 8 6
40 190.3 456.8 427.4 4.14 10.83 10.0 22.65 3.761 84.6 13.4 54.28 26.6 51.4 62.4
9 1 4 8 2 6 7 5 9
55 200.4 462.8 435.2 3.2 8 7.47 30.5 5.145 84.9 22.3 127.5 63.1 84.5 75.8
7 0 6 5 5 0 1 3 2 9
10 203.1 438.7 327.8 17.8 37.05 30.3 3.82 0.821 73.9 0.41 0.71 0.29 1.73 9.26
CF 3 2 5 2
25 214 440.4 340.5 7.6 15.36 12.6 9.42 2.065 74.3 2.26 9.81 4.12 9.64 20.5
1 9 4 3 7
40 219 442.4 348.4 4.86 9.84 8.1 15.63 3.289 74.9 5.68 39.70 16.5 24.2 32.5
1 8 9 4 3 9
55 222.4 452.2 353.0 3.6 7.35 5.97 21.28 4.615 75.5 10.4 98.93 40.6 43.8 43.0
4 8 1 5 5 9 5 8 1
452 Dv: 10-6 %K-3.min-1, Di: 10-2 %min-3, Db: 10-4 %min-4, S: 10-7 %K-3.min-2, C: 10-5 %K-2.min-1

453 In general, the main product of the pyrolysis process is the bio-oil, which
454 depends on the various parameters for the easier formation such as; short vapor residence
455 time, adequate temperate range, and higher heating rate program [37]. Taking into account

20
456 that the values of heating rate and temperature ranges significantly affect the pyrolysis
457 process. However, determining a consistent temperature and a higher heating rate under N2
458 atmosphere might be favorable to the maximum yield of bio-oil. It ought to be noted that,
459 increase in the heating rate can enhance the performance of the pyrolysis process of both non-
460 edible oilseeds up to a certain extent.

461 3.4. Kinetic parameters analysis by Model-free methods

462 Model-free isoconversioanl methods are utilized to determine the activation energy (Ea) by
463 KAS, OFW, STR, FRM, LTA, and VYZ, reaction mechanism [g(α)] by MP, pre-exponential
464 factor (A) by KN, and reaction order (n) by AVM. The detailed studies about the kinetic

of
465 parameters are discussed below:

ro
466 3.4.1. Activation energy analysis

467
468
-p
Different isoconversional methods such as some linear integrals methods (KAS, OFW, STR,
and LTA), linear differential method (FRM), and non-linear integral method (VYZ) were
re
469 utilized to estimate the activation energy (Ea) as a function of conversion (α). The linear
lP

470 curves found by the above methods are shown in Fig. S1 and S2 for both the oilseeds. The
471 consistent quality of the linear regressions found by the above methods was evaluated by
na

472 coefficients of determination (R2) in the pyrolysis process. According to all linear fitting
473 models, the conversion ranges from 0.1 to 0.8 with a step of 0.05 are chosen to evaluate the
ur

474 activation energy. It was observed from Table S1 that higher the conversion (>0.8), the data
Jo

475 did not fit well in the above model due to low linear coefficients of determination (R2). The
476 estimation of the values of activation energy (Ea) along with the degree of conversion (α)
477 evaluated through the above methods are represented in Fig. 3(a) and (b) for PR, and CF,
478 respectively.

21
of
ro
479
-p
re
480 Fig. 3(a). Variation of activation energy (Ea) as a function of conversion rate (α) by different
lP

481 model-free isoconversional methods during the reaction of PR


482
na
ur
Jo

483
484 Fig. 3(b). Variation of activation energy (Ea) as a function of conversion rate (α) by different
485 model-free isoconversional methods during the reaction of CF

22
486 In the pyrolysis process of PR, the activation energy obtained by KAS, OFW, STR, and VYZ
487 methods remains constant in the conversion ranges between 0.25 and 0.75. The differences in
488 between highest and lowest activation energy as a function of conversion (Eα) for PR
489 pyrolysis process were: 59.99% (150.44 – 240.69 kJ/mol), 58.52% (151.36 – 239.94 kJ/mol),
490 59.82% (150.77 – 240.97 kJ/mol), 89.17% (159.40 – 301.55 kJ/mol), 110.50% (143.68 –
491 302.46 kJ/mol), and 59.92% (150.66 – 240.94 kJ/mol) for KAS, OFW, STR, FRM, LTA and
492 VYZ methods, respectively. The standard deviation of Eα obtained by the VYZ method is
493 21.49 kJ/mol, which is around 10.4% of the average value of Eα (206.59 kJ/mol) with a
494 standard error of 5.5 kJ/mol.

495 Unlikely, in the CF pyrolysis process, the values of activation energy obtained

of
496 by all methods were progressively increased with an increase in the degree of conversion.

ro
497 The differences in between highest and lowest activation energy with degree of conversion
498
499
-p
(Eα) for CF pyrolysis process were: 289.23% (73.83 – 287.37 kJ/mol), 265.66% (77.65 –
283.94 kJ/mol), 286.65% (74.34 – 287.44 kJ/mol), 186.77% (116.11 – 332.98 kJ/mol),
re
500 190.94% (111.35 – 323.97 kJ/mol), 288.73% (74.14 – 288.21 kJ/mol) for KAS, OFW, STR,
lP

501 FRM, LTA and VYZ methods respectively. The average value of the Eα of CF via the VYZ
502 method is 180.35 kJ/mol, along with a standard deviation of 50.07 kJ/mol, which is nearly
na

503 27.76% of the average value followed by the standard error of 12.93 kJ/mol. It reveals that
504 the differences between them were knowingly higher than those who come across in the
ur

505 pyrolysis process of PR, and a constant range of Eα could not found on account of the CF
Jo

506 pyrolysis process.

507 From the Table 4, it was found that the mean of error of activation energy (Ea)
508 which was evaluated for several linear methods based on non-linear method (VYZ) results
509 were varied from 0.03% to 13.17% and increased in the order of STR < KAS < OFW < FRM
510 < LTA for both the oilseeds. It was also observed that the variation range of activation energy
511 (Ea) of the VYZ method is around 45 kJ/mol and 105 kJ/mol of PR and CF, respectively. In
512 the case of CF, it was the quite high value of variation range as compared to PR, which
513 indicates the average value of activation energy of CF pyrolysis cannot be used to signify the
514 variation range of Eα statistically. It was also observed that mean error (%) of the linear non-
515 approximation integral method (LTA) has maximum as compared to a linear non-
516 approximation differential method (FRM). The evaluated average activation energy values
517 are in acceptable concurrence with other studies such as; Samanea saman seeds [28], rice
518 husk, and pinewood [8], waste sawdust [33].

23
519 Table 4. The activation energy (Ea) estimates, according to Model-free Isoconversional
520 methods (KAS, OFW, STR, LTA, FRM, and VYZ).

PR
Conve KAS OFW STR FRM LTA VYZ
rsion
Ea A Err Ea A Err Ea A Err Ea A Err Ea A Err Ea
(α)
(kJ/ (min- or (kJ/ (min- or (kJ/ (min- or (kJ/ (min- or (kJ/ (min- or (kJ/
1 1 1 1 1
mol) ) (% mol) ) (% mol) ) (% mol) ) (% mol) ) (% mol)
) ) ) ) )
0.1 150. 1.23E 0.1 151. 1.60E 0.4 150. 5.01E 0.0 159. 2.61E 5.8 143. 3.22E 4.6 150.
44 +11 4 36 +17 6 77 +12 7 4 +13 0 68 +06 3 66
0.15 170. 3.25E 0.1 170. 3.64E 0.0 170. 8.66E 0.0 185. 2.80E 8.4 166. 1.80E 2.5 170.
39 +12 2 71 +18 6 68 +13 5 04 +15 6 26 +07 4 6
0.2 189. 6.76E 0.1 188. 6.60E 0.2 189. 1.32E 0.0 206. 1.14E 9.2 231. 1.28E 22. 189.
21 +13 1 91 +19 7 46 +15 2 99 +17 8 83 +09 39 42
0.25 212. 3.15E 0.1 211. 2.60E 0.6 212. 4.79E 0.0 230. 4.89E 8.3 230. 6.57E 8.5 212.
2 +15 0 03 +21 5 39 +16 1 04 +18 0 51 +09 2 41
0.3 219. 4.93E 0.0 217. 4.02E 0.6 219. 6.08E 0.0 225. 7.23E 2.6 241. 7.01E 10. 219.
16 +15 9 89 +21 7 35 +16 0 13 +17 3 75 +10 21 36

of
0.35 213. 7.07E 0.1 212. 6.33E 0.4 213. 7.26E 0.0 210. 2.07E 1.6 242. 4.13E 13. 213.
74 +14 1 97 +20 7 98 +15 1 45 +16 4 45 +11 31 97

ro
0.4 214. 4.30E 0.1 213. 3.97E 0.4 214. 3.73E 0.0 205. 5.36E 4.3 199. 7.98E 7.0 214.
62 +14 1 99 +20 0 88 +15 1 48 +15 6 67 +10 7 85
0.45 211. 1.28E 0.1 210. 1.26E 0.2 211. 9.51E 0.0 206. 4.94E 2.0 192. 1.87E 8.8 211.
08 +14 2 79 +20 6 37 +14 2 94 +15 8 7 +11 1 33
0.5

0.55
209.
9
211.
6.30E
+13
5.53E
0.1
2
0.1
209.
82
211.
6.46E
+19
5.75E
0.1
6
0.1
210.
2
212.
-p
4.06E
+14
3.11E
0.0
2
0.0
219.
51
210.
3.11E
+16
3.64E
4.4
5
0.7
220.
54
223.
1.04E
+13
5.61E
4.9
4
5.2
210.
15
212.
re
78 +13 2 76 +19 3 1 +14 3 53 +15 1 12 +13 3 04
0.6 212. 3.69E 0.1 212. 3.93E 0.0 212. 1.81E 0.0 203. 7.64E 4.2 192. 8.41E 9.3 212.
02 +13 2 14 +19 7 34 +14 3 35 +14 1 45 +12 4 28
lP

0.65 212. 3.00E 0.1 213. 3.26E 0.0 213. 1.29E 0.0 209. 1.76E 1.8 182. 9.53E 14. 213.
86 +13 3 09 +19 2 21 +14 4 1 +15 9 29 +12 47 13
0.7 212. 1.91E 0.1 212. 2.13E 0.0 212. 7.16E 0.0 219. 9.32E 3.5 200. 3.34E 5.5 212.
07 +13 3 46 +19 5 43 +13 4 93 +15 7 49 +14 8 35
na

0.75 215. 2.42E 0.1 215. 2.69E 0.0 215. 7.92E 0.0 239. 2.02E 11. 244. 4.34E 13. 215.
12 +13 3 48 +19 4 47 +13 4 95 +17 40 66 +17 59 39
0.8 240. 1.43E 0.1 239. 1.31E 0.4 240. 4.03E 0.0 301. 3.34E 25. 302. 5.05E 25. 240.
69 +15 0 94 +21 1 97 +15 1 55 +21 16 46 +21 53 94
ur

Avg. 206. 7.38E 0.1 206. 6.27E 0.2 206. 8.49E 0.0 215. 2.23E 6.2 214. 3.36E 10. 206.
35 ± +14 2 16 ± +20 7 64 ± +15 3 56 ± +20 6 32 ± +20 41 59 ±
5.55 5.45 5.55 7.84 9.90 5.55
Jo

CF
KAS OFW STR FRM LTA VYZ
0.1 73.8 1.35E 0.4 77.6 5.61E 4.7 74.3 5.41E 0.2 116. 1.40E 56. 111. 1.51E 50. 74.1
3 +04 0 5 +10 4 4 +05 8 11 +10 61 35 +06 20 4
0.15 134. 1.83E 0.1 135. 2.73E 0.8 134. 4.78E 0.0 149. 3.52E 10. 193. 1.07E 43. 134.
1 +10 7 46 +16 4 45 +11 8 00 +13 91 43 +08 99 34
0.2 144. 1.09E 0.1 145. 1.49E 0.5 144. 2.09E 0.0 154. 6.00E 6.9 177. 2.53E 23. 144.
26 +11 6 34 +17 9 59 +12 7 55 +13 6 80 +08 05 49
0.25 154. 5.69E 0.1 155. 7.21E 0.3 155. 8.51E 0.0 162. 1.54E 4.8 165. 4.35E 6.6 154.
73 +11 5 55 +17 8 05 +12 6 46 +14 4 24 +08 4 96
0.3 160. 9.29E 0.1 161. 1.16E 0.3 160. 1.13E 0.0 169. 3.72E 5.3 178. 3.80E 10. 160.
35 +11 4 12 +18 4 67 +13 6 2 +14 7 01 +09 86 58
0.35 164. 1.29E 0.1 165. 1.60E 0.3 164. 1.30E 0.0 170. 3.57E 3.6 178. 1.65E 8.2 164.
64 +12 4 39 +18 1 96 +13 5 94 +14 8 53 +10 8 87
0.4 169. 2.08E 0.1 169. 2.53E 0.2 169. 1.77E 0.0 176. 8.10E 4.3 186. 1.33E 9.7 169.
26 +12 4 95 +18 7 58 +13 5 8 +14 1 03 +11 6 49
0.45 172. 2.80E 0.1 173. 3.38E 0.2 173. 2.05E 0.0 181. 1.25E 4.5 195. 1.44E 12. 173.
96 +12 4 63 +18 5 28 +13 5 03 +15 3 39 +12 82 19
0.5 179. 7.31E 0.1 180. 8.45E 0.1 180. 4.63E 0.0 188. 3.46E 4.6 190. 3.59E 5.7 180.
77 +12 3 26 +18 4 09 +13 5 3 +15 1 32 +12 3 00
0.55 186. 1.82E 0.1 186. 2.02E 0.0 186. 1.00E 0.0 196. 1.02E 5.0 192. 1.79E 3.1 186.
51 +13 3 82 +19 4 82 +14 4 12 +16 2 66 +13 7 74
0.6 194. 5.86E 0.1 194. 6.17E 0.0 195. 2.82E 0.0 205. 4.15E 5.4 193. 7.43E 0.9 194.
73 +13 2 79 +19 9 03 +14 3 68 +16 9 18 +13 1 97
0.65 207. 4.35E 0.1 207. 4.35E 0.3 207. 1.83E 0.0 231. 2.84E 11. 205. 1.34E 0.9 207.
54 +14 1 14 +20 0 81 +15 2 01 +18 19 81 +15 4 77
0.7 225. 6.52E 0.1 224. 5.58E 0.5 225. 2.41E 0.0 250. 3.61E 11. 242. 5.13E 7.3 225.

24
46 +15 1 4 +21 8 69 +16 1 92 +19 17 27 +17 4 71
0.75 245. 8.87E 0.1 243. 6.82E 0.8 245. 2.86E 0.0 260. 4.54E 5.8 249. 3.19E 1.4 245.
35 +16 5 6 +22 6 53 +17 8 04 +19 2 17 +18 0 73
0.8 287. 3.82E 0.2 283. 2.32E 1.4 287. 1.07E 0.2 332. 2.89E 15. 323. 3.12E 12. 288.
37 +19 9 94 +25 8 44 +20 7 98 +24 54 97 +23 41 21
Avg. 180. 2.55E 0.1 180. 1.55E 0.7 180. 7.15E 0.0 196. 1.93E 10. 198. 2.08E 13. 180.
06 ± +18 7 34 ± +24 5 36 ± +18 8 34 ± +23 40 88 ± +22 17 35 ±
12.9 12.5 12.8 13.9 12.0 12.9
0 0 8 1 6 3
L¤¥¦ L§•¨,©ª«,¨DC,ªC¬,-D•
z::(: % = £ £ × 100
L¤¥¦
521

522 3.4.2. Reaction mechanism analysis

523 The reaction mechanism defines the order of straightforward reactions, which must happen to
524 carry forward from reactants to products. In order to evaluate the reaction mechanism, the
525 average value of activation energy obtained through a non-linear model-free method (VYZ)

of
526 was considered. This method is more precise when the variation range of activation energy is

ro
527 low, and the activation energy as a function of conversion throughout the process is constant.
528 Nevertheless, this was done only to predict the selectness of the reaction model for the
529 pyrolysis of PR and CF.
-p
re
530 The experimental master-plots p(x)α/p(x)0.5 and theoretical master-plots
lP

531 g(α)/g(0.5) along with the degree of conversion (α) of PR and CF are represented in Fig. 4(a)
532 and (b) respectively. The values of theoretical master-plots were obtained from all reaction
na

533 models according to Table – 1, and the values of experimental master-plots were determined
534 based on the 8th order Senum-Yang’s approximation equation (27) at 10 oC/min heating rate.
ur

535 In this method, the predicted reaction model has to be determined when the experimental
Jo

536 master-plots were overlapped with theoretical master-plots.

25
of
ro
537 -p
re
538 Fig. 4 (a). Comparison between theoretical and experimental master plots under a heating
539 rate of 10 oC/min based on 21 different kinetic models by using Criado-Master plots method
of PR
lP

540
541
na
ur
Jo

542

543 Fig. 4 (b). Comparison between theoretical and experimental master plots under a heating
544 rate of 10 oC/min based on 21 different kinetic models by using Criado-Master plots method
545 of CF
26
546 From Fig. 4(a), it was noticed that during the PR pyrolysis process, the
547 Power-law model is the most consistent due to similar trends has occurred of the
548 experimental data as P3 for conversion (α) range from 0.05 to 0.85; also the Nucleation
549 model (A4) was well fitted when conversion (α) was in the range of 0.2 – 0.65. Similarly,
550 from Fig. 4(b) for the CF pyrolysis process, the Power-law and Nucleation model has the
551 same trends as P3 (when α = 0.15 – 0.6), and A4 (when α = 0.3 – 0.7) respectively. Similar
552 observation was obtained in detailed assessment of pyrolysis kinetics of invasive
553 lignocellulosic biomasses (Prosopis juliflora and Lantana camara) [9].

554 Similarly, Dhyani et al., 2017 reported that the reaction mechanism of
555 sorghum straw pyrolysis was nucleation and growth mechanism followed by A2, A3, and A4

of
556 [38]. The reaction mechanisms found by Kuang et al., 2019 were F1, A3, and A4 for kerogen

ro
557 from the Green River oil shale [39]. The power-law mechanism prefers to used basically for
558
559
-p
heterogeneous reactions [40]; generally, biomass feedstocks are also heterogeneous in nature.
Therefore, during the pyrolysis of biomass, the heterogeneous (gas-solid) reactions of biochar
re
560 and the secondary gas-phase reaction of gas and tar were released [41], therefore these
lP

561 feedstocks were well fitted with the power-law mechanism.

562 3.4.3. Pre-exponential factor analysis


na

563 The pre-exponential factor (A) for both the biomasses obtained from Kissinger’s Equation
ur

564 using the Eα values from the non-linear integral method (VYZ) and the peak temperature of
565 the DTG curve of 10 oC/min heating rate from Table – 6. The values of A were varied from
Jo

566 3.42E+06 s-1 to 5.67E+13 s-1 and 1.17E+01 s-1 to 1.94E+20 s-1 for PR and CF respectively.
567 The variation of A values (101 – 1020 s-1) from 0.1 – 0.8 signified the complex reaction
568 happened during the thermal degradation of materials [28]. The mean value of the A was
569 4.08E+12 s-1 and 1.29E+19 s-1 of PR and CF, respectively. The value of the A depends on the
570 temperature due to the molecular collision of reactants [42]. The higher value of the A
571 attributed to a simple activated complex [43]. In the present study, it was found that CF had
572 higher value of A than PR.

573 3.5. Reaction order analysis

574 The order of reaction (n) signifies the reaction swiftness with respect to time. Various
575 chemical processes were involved during the pyrolysis of biomass to assess the order of
576 reaction based on temperature. Here, fifteen levels of temperature (from 150 oC to 500 oC)
577 were additionally utilized at four different heating rate programs of 10, 25, 40, and 55

27
o
578 C/min. Implementing the Avrami theory, the linear fit regression plots of PR and CF are
579 presented in Fig. S3. The values of reaction order from the slope of the linear equations
580 followed by the coefficient of determination (R2) are tabulated in Table - S2 for both the
581 oilseeds. It was noticed that the reaction orders of both the oilseeds were varied significantly
582 instead of remained constant due to the degradation of hemicellulose, cellulose, and lignin at
583 different temperature ranges. An expeditious increase in reaction order (n) from 0.2883 (150
o
584 C) to 0.4917 (175 oC) for PR signifies that the light volatiles compound and hemicellulose
585 degradation occurs very fast at this stage due to rapid change of order of the reaction.
586 Similarly, in case of CF, the reaction order (n) remains constant from 150 oC (n = 0.1676) to
587 200 oC (n = 0.1583), and after that it was increased instantly up to 0.3024 (225 oC). From the

of
588 Table - S2, it was also noticed that the reaction order (n) of PR pyrolysis first increased from
0.2883 to 0.4917, then decreased to 0.1624 (325 oC), then again increased to 0.2731 (425 oC)

ro
589
590 and finally decreased to 0.025 (500 oC). Unlikely in the CF pyrolysis process, the order of
591
-p
reaction first increased from 0.1583 to 0.3024, and then steadily decreased to 0.0407 (500
re
o
592 C). Similar results was also found in olive mill solid waste (OMSW) pyrolysis, the reaction
593 order first increased from 0.1004 (517K) to 0.1787 (644K), and then it was decreased to
lP

594 0.1220 (738K) [24]. Therefore, it can be noted that varying the values of activation energy
595 with conversion and reaction order with temperature for PR and CF are different due to
na

596 alterations in their composition and complexity in their nature.


ur

597 3.6. Kinetic parameters analysis by Model-fitting method


Jo

598 In CR method, the model-fitting method works on single heating rate. This model has been
599 used widely to estimate the reaction order and reaction mechanism based on the activation
600 energy (Ea) and co-efficient of determination (R2). In the present work, the CR method was
601 used at various heating rates (β = 10, 25, 40, and 55 oC/min). It is important to mention that
602 the values of reaction orders and pre-exponential factor determined from the above model-
603 free isoconversional methods, have no substantial dominance; it was only measured by way
604 of fitting parameters [25]. The reaction order was determined using the trial and error
605 method on the CR model by putting the random values of reaction order (n) such as n = 0.1,
606 0.2, 0.3, ……. ∞ at different heating rates. The best-fitted values of reaction order of PR were
607 determined as n = 0.09, 0.4, 0.55, and 0.5 at β = 10, 25, 40, and 55 oC/min respectively.
608 Similarly, in case of CF, the values of reaction order were found to be n = 0.8, 0.95, 1.2, and
609 1.1 while β = 10, 25, 40, and 55 oC/min respectively. From the Table - S3, it is observed that
610 the similar trends of reaction order along with heating rates were carried out for both the

28
611 oilseeds and the reaction order is directly proportional to the heating rate up to some extent.
612 While N = 1, the values of activation energy (Ea) and pre-exponential factor (A) of PR were
613 28.33 kJ/mol, 29.67 kJ/mol, 30.60 kJ/mol, and 31.66 kJ/mol along with 7.52E+02 s-1,
614 2.15E+03 s-1, 3.99E+03 s-1, and 6.41E+03 s-1 at 10, 25, 40, and 55 oC/min heating rate
615 respectively. Similarly, the CF has same trends of activation energy (Ea) and pre-exponential
616 factor (A) values with increasing of heating rates such as 28.09 kJ/mol, 30.33 kJ/mol, 30.73
617 kJ/mol, and 31.08 kJ/mol followed by 1.50E+03 s-1 5.55E+03 s-1, 8.77E+03 s-1, and
618 1.18E+04 s-1 for 10, 25, 40, and 55 oC/min at N = 1. When N ≠ 1, the values of activation
619 energy (Ea) were determined to be 23.46 kJ/mol, 26.32 kJ/mol, 27.42 kJ/mol, and 28.85
620 kJ/mol, whereas the pre-exponential factors were 1.85E+02 s-1, 8.44E+02 s-1, 1.66E+03 s-1,

of
621 and 3.00E+03 s-1 for PR at 10, 25, 40, and 55 oC/min heating rate respectively. On the other

ro
622 hand, the activation energy and pre-exponential factor as a function of heating rate at N ≠ 1
623 for CF were found to be 27.71, 28.69, 32.40, and 31.88 kJ/mol followed by 1.34E+03,
624
-p
3.50E+03, 1.39E+04, and 1.47E+04 s-1 of 10, 25, 40, and 55 oC/min respectively. The
re
625 coefficient of determination (R2) was > 0.99 for all the above cases, which implied the best
626 fitting. Hence, it was observed that the values of activation energy (Ea) and pre-exponential
lP

627 factor (A) are directly proportional to the heating rate. Pyrolysis is an unremitting process
628 that implies an increase in heating rates and temperatures; the volatilization increases,
na

629 employing significant heat transfer among particles of biomass [9].


ur

630 In solid-state reactions, various methods and models have been used to
Jo

631 determine the appropriate reaction mechanism dependent on certain mechanistic assumptions.
632 A proper reaction model was chosen based on the compared the value of activation energy
633 (Ea) obtained by model-free isoconversional method (VYZ) and also the highest linear fit
634 coefficient (R2). Table – 1 shows five groups of theoretical (both integral and differential)
635 reaction models, such as geometrical contraction model (Rn), diffusional (Dn), nucleation
636 (An), an order based (Fn), and power-law (Pn). The values of activation energy and co-
637 efficient of determination were tabulated in Table – 5 for each of the reaction models at 10
o
638 C/min heating rate. It can be observed that the reaction models (Rn, Dn, and Pn) did not
639 exhibit the excellent value of R2 for both the biomasses. However, in the case of PR, the
640 nucleation (An) and order-based (Fn) models show good agreement in co-efficient of
641 determination (R2 > 0.99). Similarly, in the case of CF, only order-based (Fn) model was well
642 fitted with the experimental thermal data followed by the highest value of R2 > 0.99. From
643 Table – 5, it can also be noticed that the order-based (F2) model was shown to have the

29
644 highest coefficient of determination (R2) and appropriately match with the nearest average
645 value of activation energy (Ea) determined from isoconversional method (VYZ) as listed in
646 Table – 4. The reaction mechanism of CF was probably the order-based (F3) model due to the
647 similar value of activation energy obtained by the VYZ method and the highest value of R2
648 among all of the models presented in Table – 5.

649 Table 5. Results of Reaction mechanism by Model-fitting Isoconversioanl method (Coats-


650 Redfern method)
PR CF
2
Reaction Model Ea (kJ/mol) R Ea (kJ/mol) R2
R2 162.12 0.9847 151.31 0.9778
R3 177.09 0.9879 167.11 0.9812

of
D1 267.71 0.973 246.93 0.9675
D2 298.47 0.9805 276.02 0.9736

ro
D3 354.18 0.9879 334.22 0.9812
D4 345.86 0.985 327.57 0.9793
A1.5
A2
102.26
76.49
-p
0.9902
0.9902
91.45
68.17
0.9789
0.9789
re
A3 50.72 0.9902 45.73 0.9789
A4/5 191.22 0.9902 170.44 0.9789
A5/6 183.74 0.9902 163.79 0.9789
lP

F1 152.98 0.9902 136.35 0.9789


F1.5 164.62 0.9957 145.50 0.9844
na

F2 178.75 0.9976 154.64 0.9886


F3 209.51 0.9904 177.92 0.9912
F4 245.26 0.9739 203.69 0.9868
ur

P1.5 201.20 0.973 184.57 0.9675


P2 67.34 0.973 61.52 0.9675
Jo

P3 44.90 0.973 40.74 0.9675


P4 33.26 0.973 30.76 0.9675
651

652

653 In the model-fitting method (CR model), the predictions of the model depend
654 upon the linear fit coefficient (R2). If the coefficient of determination (R2) would be high for
655 a model, then most probably, that model would be likely to choose the best fit for that
656 particular process. From this point of view, the credibility of the model-fitting method is not
657 good enough to determine the kinetics assessments of reaction for a particular process.
658 Generally, biomass pyrolysis is a complex process and very difficult to determine the
659 appropriate reaction mechanism throughout the process. Besides, it is essential to use of
660 Master-plot method (MP) to validate the CR method results. The Master-plot (MP) method

30
661 by Criado comes up with more precise and well-founded info for the estimation of reaction
662 models, as highly recommended by the ICTAC and Kinetic Committee [6].

663 3.7. Thermodynamic parameters analysis

664 The thermodynamic parameters, such as Gibbs free energy (∆G), change of enthalpy (∆H),
665 and change of entropy (∆S), were determined by utilizing equations (41), (42), and (43)
666 respectively. The activation energy as a function of conversion (Eα) obtained from the model-
667 free isoconversional method (VYZ) was utilized to estimate the thermodynamic parameters.
668 The thermodynamic analysis of PR and CF is tabulated in Table – 6.

669 Table 6. Results of Thermodynamic parameters of PR and CF

of
PR

ro
Conversion A (s-1) ∆H (kJ/mol) ∆G (kJ/mol) ∆S (J/mol)
0.1 3.42E+06 145.07 235.75 -134.91
0.15 1.37E+08 165.01 235.05 -104.21
0.2
0.25
4.42E+09
3.03E+11
-p
183.83
206.82
234.47
233.83
-75.34
-40.18
re
0.3 1.09E+12 213.77 233.65 -29.57
0.35 4.04E+11 208.38 233.79 -37.80
0.4 4.75E+11 209.26 233.76 -36.45
lP

0.45 2.49E+11 205.74 233.86 -41.83


0.5 2.00E+11 204.56 233.89 -43.63
0.55 2.83E+11 206.45 233.84 -40.74
na

0.6 2.96E+11 206.69 233.83 -40.38


0.65 3.46E+11 207.54 233.81 -39.08
0.7 3.00E+11 206.76 233.83 -40.27
ur

0.75 5.24E+11 209.80 233.75 -35.63


0.8 5.67E+13 235.35 233.12 3.31
Avg. 4.08E+12 201.00 234.02 -49.11
Jo

CF
0.1 1.17E+01 69.14 212.36 -238.61
0.15 3.67E+06 129.35 209.39 -133.36
0.2 3.03E+07 139.50 209.03 -115.83
0.25 2.64E+08 149.97 208.68 -97.82
0.3 8.45E+08 155.59 208.50 -88.15
0.35 2.05E+09 159.88 208.37 -80.78
0.4 5.32E+09 164.50 208.23 -72.86
0.45 1.14E+10 168.20 208.12 -66.51
0.5 4.64E+10 175.01 207.93 -54.85
0.55 1.86E+11 181.75 207.75 -43.31
0.6 1.01E+12 189.98 207.53 -29.25
0.65 1.40E+13 202.78 207.21 -7.39
0.7 5.53E+14 220.72 206.80 23.19
0.75 3.33E+16 240.74 206.38 57.25
0.8 1.94E+20 283.21 205.58 129.35
Avg. 1.29E+19 175.35 208.12 -54.60
670 The value of ∆G reflects the overall energy increased of the system at the perspectives of
671 reagents and the origination of the activated product [13]. The ∆G is directly related to the
672 other parameters, such as ∆H and ∆S of the formation of the products. The value of ∆G
673 varied from 235.75 kJ/mol to 233.12 kJ/mol and 212.36 kJ/mol to 205.58 kJ/mol for PR and
31
674 CF respectively. Meanwhile, the value of change in enthalpy (∆H) signifies that the
675 difference of energy between the reagents and the formation of decomposition products [44].
676 A small potential barrier (Ea – ∆H) was observed during the pyrolysis of PR (< 5.56 kJ/mol)
677 and CF (< 5.00 kJ/mol), which indicates that a little amount of additional energy is needed to
678 achieve the formation of final products. The value of change of entropy (∆S) was increased
679 while the conversion is increased, and it alters from most -ve values to +ve values. In the
680 present study, the ∆S values were altered from -134.91 J/mol-K to 3.31 J/mol-K and -238.61
681 J/mol-K to 129.35 J/mol-K for PR and CF, respectively. The -ve value of ∆S signifies that
682 the formation of products has a lower degree of disorders than biomass, while +ve value
683 implies the high reactivity, i.e., the system can react quicker to the origination of the activated

of
684 complex. The +ve value of ∆G, ∆H, and ∆S indicates a non-spontaneous reaction has

ro
685 occurred and dependent on the heat introduced to the process during the pyrolysis of PR and
686 CF. Similar results were also found for Samanea saman seeds [28] and Brazilian pine-fruit
687 shell [45].
-p
re
688 4. Conclusion
lP

689 Kinetic parameters [Ea, A, g(α), and n] and thermodynamic parameters of Putranjiva
690 roxburghii (putranjiva) (PR) and Cassia fistula (amaltas) (CF) were investigated through
na

691 non-isothermal TGA under N2 atmosphere using both model-fitting and model-free
692 approaches. The physicochemical characterization and pyrolysis performance indices (PPI)
ur

693 confirmed that PR produced more volatiles as compared to CF. However, from the kinetic
Jo

694 analysis, it was noticed that CF has lower activation energy, higher pre-exponential factor,
695 and higher reaction order than PR. The apparent activation energy (Eα) values calculated
696 from different isoconversional methods displayed a variable value according to their
697 approximation equation. The thermodynamic parameters indicated that the CF has lower
698 values of ∆G, ∆H, and ∆S than PR. The results confirmed that these two oilseeds (PR and
699 CF) can be used as potential feedstock to produce good quality fuels and value-added
700 chemicals. Further, the kinetic and thermodynamic parameters estimated and reported in this
701 work will be of great importance to scale-up and optimizing pyrolysis processes for such
702 feedstocks.

703 Acknowledgments:

704 Authors would like to thank the Analytical Facility, Department of Chemical
705 Engineering, Indian Institute of Technology Guwahati and Centre of Excellence – Green &

32
706 Efficient Energy Technology (CoE-GEET), CUJ, Ranchi for financial and other necessary
707 supports for carrying out this research work.

708 Appendix A. Supplementary data

709 E-supplementary data for this work can be found in e-version of this paper online.

710 References

711 [1] S. Clarke, F. Preto, Biomass Burn Characteristics, Minist. Agric. Food Rural Aff.
712 (2011). http://www.omafra.gov.on.ca/english/engineer/facts/11-033.pdf (accessed 08
713 April 2020).

of
714 [2] A. Kumar, N. Kumar, P. Baredar, A. Shukla, A review on biomass energy resources,
potential, conversion and policy in India, Renew. Sustain. Energy. Rev. (2015).

ro
715
716 https://doi.org/10.1016/j.rser.2015.02.007.

717 [3].
-p
A. Sahoo, S. Kumar, K. Mohanty, A comprehensive characterization of non-edible
re
718 lignocellulosic biomass to elucidate their biofuel production potential. Biomass. Conv.
lP

719 Bioref. (2020) https://doi.org/10.1007/s13399-020-00924-6.

720 [4] Y.W. Huang, M.Q. Chen, Y. Li, An innovative evaluation method for kinetic
na

721 parameters in distributed activation energy model and its application in


722 thermochemical process of solid fuels, Thermochim. Acta. (2017).
ur

723 https://doi.org/10.1016/j.tca.2017.06.009.
Jo

724 [5] T.K. Vo, H.V. Ly, O.K. Lee, E.Y. Lee, C.H. Kim, J.W. Seo, J. Kim, S.S. Kim,
725 Pyrolysis characteristics and kinetics of microalgal Aurantiochytrium sp. KRS101,
726 Energy. (2017). https://doi.org/10.1016/j.energy.2016.12.040.

727 [6] S. Vyazovkin, A.K. Burnham, J.M. Criado, L.A. Pérez-Maqueda, C. Popescu, N.
728 Sbirrazzuoli, ICTAC Kinetics Committee recommendations for performing kinetic
729 computations on thermal analysis data, Thermochim. Acta. (2011).
730 https://doi.org/10.1016/j.tca.2011.03.034.

731 [7] Q. He, L. Ding, Y. Gong, W. Li, J. Wei, G. Yu, Effect of torrefaction on pinewood
732 pyrolysis kinetics and thermal behavior using thermogravimetric analysis, Bioresour.
733 Technol. (2019). https://doi.org/10.1016/j.biortech.2019.01.138.

734 [8] C.N. Arenas, M.V. Navarro, J.D. Martínez, Pyrolysis kinetics of biomass wastes using

33
735 isoconversional methods and the distributed activation energy model, Bioresour.
736 Technol. (2019). https://doi.org/10.1016/j.biortech.2019.121485.

737 [9]. A. Sahoo, S. Kumar, J. Kumar, T. Bhaskar, A detailed assessment of pyrolysis kinetics
738 of invasive lignocellulosic biomasses (Prosopis juliflora and Lantana camara) by
739 thermogravimetric analysis. Bioresour. Technol. 319 (2021) 124060.
740 https://doi.org/10.1016/j.biortech.2020.124060.

741 [10]. R.K. Mishra, V. Kumar, K. Mohanty, Pyrolysis kinetics behaviour and thermal
742 pyrolysis of Samanea saman seeds towards the production of renewable fuel. J. Energ.
743 Inst. 93 (2020) 1148-1162. https://doi.org/10.1016/j.joei.2019.10.008.

of
744 [11]. V.O. Santos, L.S. Queiroz, R.O. Araujo, F.C.P. Ribeiro, M.N. Guimaraes, C.E.F. da

ro
745 Costa, J.S. Chaar, L.K.C. de Souza, Pyrolysis of acai seed biomass: Kinetics and
746 thermodynamic parameters using thermogravimetric analysis. Bioresour. Technol.
747
-p
Rep. 12 (2020) 100553. https://doi.org/10.1016/j.biteb.2020.100553.
re
748 [12] W. Xie, S. Wen, J. Liu, W. Xie, J. Kuo, X. Lu, S. Sun, K. Chang, M. Buyukada, F.
lP

749 Evrendilek, Comparative thermogravimetric analyses of co-combustion of textile


750 dyeing sludge and sugarcane bagasse in carbon dioxide/oxygen and nitrogen/oxygen
na

751 atmospheres: Thermal conversion characteristics, kinetics, and thermodynamics,


752 Bioresour. Technol. (2018). https://doi.org/10.1016/j.biortech.2018.01.110.
ur

753 [13] S. Wen, Y. Yan, J. Liu, M. Buyukada, F. Evrendilek, Pyrolysis performance, kinetic,
Jo

754 thermodynamic, product and joint optimization analyses of incense sticks in N2 and
755 CO2 atmospheres, Renew. Energy. (2019).
756 https://doi.org/10.1016/j.renene.2019.04.040.

757 [14] V. Dhyani, T. Bhaskar, Kinetic Analysis of Biomass Pyrolysis, in: Waste Biorefinery,
758 (2018). https://doi.org/10.1016/b978-0-444-63992-9.00002-1.

759 [15] P. Murray, J. White, Kinetics of the thermal dehydration of clay. Part IV.
760 Interpretation of the differential thermal analysis of the clay minerals., Trans. Br.
761 Ceram. Soc. (1955). 255-264.

762 [16] C.D. Doyle, Estimating isothermal life from thermogravimetric data, J. Appl. Polym.
763 Sci. (1962). https://doi.org/10.1002/app.1962.070062406.

764 [17] M.J. Starink, The determination of activation energy from linear heating rate

34
765 experiments: A comparison of the accuracy of isoconversion methods, Thermochim.
766 Acta. (2003). https://doi.org/10.1016/S0040-6031(03)00144-8.

767 [18] C.R. Li, T.B. Tang, A new method for analysing non-isothermal thermoanalytical data
768 from solid-state reactions, Thermochim. Acta. (1999). https://doi.org/10.1016/S0040-
769 6031(98)00568-1.

770 [19] H.L. Friedman, Kinetics of thermal degradation of char-forming plastics from
771 thermogravimetry. Application to a phenolic plastic, J. Polym. Sci. Part C Polym.
772 Symp. (2007). https://doi.org/10.1002/polc.5070060121.

773 [20] S. Vyazovkin, Advanced isoconversional method, J. Therm. Anal. (1997).

of
774 https://doi.org/10.1007/bf01983708.

ro
775 [21] L.A. Pérez-Maqueda, J.M. Criado, Accuracy of Senum and Yang’s approximations to
776
777 https://doi.org/10.1023/A:1010115926340.
-p
the Arrhenius integral, J. Therm. Anal. Calorim. (2000).
re
778 [22] P.E. Sánchez-Jiménez, L.A. Pérez-Maqueda, A. Perejón, J.M. Criado, Generalized
lP

779 master plots as a straightforward approach for determining the kinetic model: The case
780 of cellulose pyrolysis, Thermochim. Acta. (2013).
na

781 https://doi.org/10.1016/j.tca.2012.11.003.
ur

782 [23] H.E. Kissinger, Variation of peak temperature with heating rate in differential thermal
Jo

783 analysis, J. Res. Natl. Bur. Stand. (1934). (1956). https://doi.org/10.6028/jres.057.026.

784 [24] M.Y. Guida, H. Bouaik, L. El Mouden, A. Moubarik, A. Aboulkas, K. El harfi, A.


785 Hannioui, Utilization of Starink Approach and Avrami Theory to Evaluate the Kinetic
786 Parameters of the Pyrolysis of Olive Mill Solid Waste and Olive Mill Wastewater, J.
787 Adv. Chem. Eng. (2017). https://doi.org/10.4172/2090-4568.1000155.

788 [25] T. Damartzis, D. Vamvuka, S. Sfakiotakis, A. Zabaniotou, Thermal degradation


789 studies and kinetic modeling of cardoon (Cynara cardunculus) pyrolysis using
790 thermogravimetric analysis (TGA), Bioresour. Technol. (2011).
791 https://doi.org/10.1016/j.biortech.2011.02.060.

792 [26] A.W. Coats, J.P. Redfern, Kinetic parameters from thermogravimetric data, Nature.
793 (1964). https://doi.org/10.1038/201068a0.

794 [27] K. Jayaraman, M.V. Kok, I. Gokalp, Thermogravimetric and mass spectrometric (TG-

35
795 MS) analysis and kinetics of coal-biomass blends, Renew. Energy. (2017).
796 https://doi.org/10.1016/j.renene.2016.08.072.

797 [28] R.K. Mishra, A. Sahoo, K. Mohanty, Pyrolysis kinetics and synergistic effect in co-
798 pyrolysis of Samanea saman seeds and polyethylene terephthalate using
799 thermogravimetric analyser, Bioresour. Technol. 289 (2019) 121608.
800 https://doi.org/10.1016/j.biortech.2019.121608.

801 [29] K.P. Shadangi, K. Mohanty, Characterization of nonconventional oil containing seeds
802 towards the production of bio-fuel, J. Renew. Sustain. Energy. 5 (2013).
803 https://doi.org/10.1063/1.4808029.

of
804 [30] S.P. Pandey, S. Kumar, Valorization of argemone mexicana seeds to renewable fuels by

ro
805 thermochemical conversion process. J. Environ. Chem. Eng. (2020)
806 https://doi.org/10.1016/j.jece.2020.104271.

807
-p
[31] N.K. Nayan, S. Kumar, R.K. Singh, Production of the liquid fuel by thermal pyrolysis
re
808 of neem seed, in: Fuel, 2013. https://doi.org/10.1016/j.fuel.2012.08.058.
lP

809 [32] G. Ozsin, A.E. Putun, Kinetics and evolved gas analysis for pyrolysis of food
810 processing wastes using TGA/MS/FT-IR, Waste Manag. (2017).
na

811 https://doi.org/10.1016/j.wasman.2017.03.020.
ur

812 [33] R.K. Mishra, K. Mohanty, Pyrolysis kinetics and thermal behavior of waste sawdust
Jo

813 biomass using thermogravimetric analysis, Bioresour. Technol. (2018).


814 https://doi.org/10.1016/j.biortech.2017.12.029.

815 [34] B. Pecha, M. Garcia-perez, Pyrolysis of Lignocellulosic Biomass: Oil, Char, and Gas,
816 In Bioenergy, 2015. https://doi.org/10.1016/B978-0-12-407909-0.00026-2.

817 [35] J.E. White, W.J. Catallo, B.L. Legendre, Biomass pyrolysis kinetics: A comparative
818 critical review with relevant agricultural residue case studies, J. Anal. Appl. Pyrolysis.
819 (2011). https://doi.org/10.1016/j.jaap.2011.01.004.

820 [36] D. Vamvuka, E. Kakaras, E. Kastanaki, P. Grammelis, Pyrolysis characteristics and


821 kinetics of biomass residuals mixtures with lignite, in: Fuel, 2003.
822 https://doi.org/10.1016/S0016-2361(03)00153-4.

823 [37] H.B. Goyal, D. Seal, R.C. Saxena, Bio-fuels from thermochemical conversion of
824 renewable resources: A review, Renew. Sustain. Energy Rev. (2008).

36
825 https://doi.org/10.1016/j.rser.2006.07.014.

826 [38] V. Dhyani, J. Kumar, T. Bhaskar, Thermal decomposition kinetics of sorghum straw
827 via thermogravimetric analysis, Bioresour. Technol. (2017).
828 https://doi.org/10.1016/j.biortech.2017.08.189.

829 [39] W. Kuang, M. Lu, I. Yeboah, G. Qian, X. Duan, J. Yang, D. Chen, X. Zhou, A
830 comprehensive kinetics study on non-isothermal pyrolysis of kerogen from Green
831 River oil shale, Chem. Eng. J. (2019). https://doi.org/10.1016/j.cej.2018.10.212.

832 [40] M. Shacham, N. Brauner, Using the power-law rate expression for assessment of rate
833 data and detection of infeasible mechanisms for reversible reactions, Ind. Eng. Chem.

of
834 Res. (1996). https://doi.org/10.1021/ie950029p.

ro
835 [41] E. Ranzi, P.E.A. Debiagi, A. Frassoldati, Mathematical Modeling of Fast Biomass
836
837
-p
Pyrolysis and Bio-Oil Formation. Note I: Kinetic Mechanism of Biomass Pyrolysis,
ACS Sustain. Chem. Eng. (2017). https://doi.org/10.1021/acssuschemeng.6b03096.
re
838 [42] J.M. Criado, L.A. Pérez-Maqueda, P.E. Sánchez-Jiménez, Dependence of the
lP

839 preexponential factor on temperature, J. Therm. Anal. Calorim. (2005).


840 https://doi.org/10.1007/s10973-005-0948-3.
na

841 [43] X. Yuan, T. He, H. Cao, Q. Yuan, Cattle manure pyrolysis process: Kinetic and
ur

842 thermodynamic analysis with isoconversional methods, Renew. Energy. (2017).


Jo

843 https://doi.org/10.1016/j.renene.2017.02.026.

844 [44] Y. Xu, B. Chen, Investigation of thermodynamic parameters in the pyrolysis


845 conversion of biomass and manure to biochars using thermogravimetric analysis,
846 Bioresour. Technol. (2013). https://doi.org/10.1016/j.biortech.2013.07.086.

847 [45] J.L.F. Alves, J.C.G. Da Silva, V.F. da Silva Filho, R.F. Alves, W.V. de Araujo
848 Galdino, S.L.F. Andersen, R.F. De Sena, Determination of the Bioenergy Potential of
849 Brazilian Pine-Fruit Shell via Pyrolysis Kinetics, Thermodynamic Study, and Evolved
850 Gas Analysis, Bioenergy Res. (2019). https://doi.org/10.1007/s12155-019-9964-1.

851

852

853

37
854 Abbreviations
855 PR: Putranjiva roxburghii
856 CF: Cassia fistula
857 TGA: Thermogravimetric analyzer
858 DTG: Differential thermogravimetric
859 PPI: Pyrolysis performance index
860 FRM: Friedman
861 KAS: Kissinger-Akahira-Sunose
862 OFW: Ozawa-Flynn-Wall

of
863 STR: Starink

ro
864 LTA: Li and Tang
865 VYZ: Vyazovkin
866 KN: Kissinger
-p
re
867 AVM: Avrami
lP

868 MP: Master plot


869 CR: Coats-Redfern
na

870 Ti: Inginition temperature


Tp: Peak temperature
ur

871

872 Tb: Burnout temperature


Jo

873 Dv: Devolatilization index


874 Di: Ignition indices
875 Db: Burnout indices
876 S: Combustion index
877 C: Flammability index
878 Wt: Total weight loss
879 α: Degree of conversion
880 Ea: Activation energy
881 Eα: Activation energy at a particular conversion
882 β: Heating rate
883 -dw/dtmax:Maximum degradation rate

38
1 List of Tables
2 Table 1. 22 kinetic models (f(α)) and their integral expression (g(α)) of the solid-state
3 pyrolysis process.
4 Table 2. Results of the physicochemical characterization of PR and CF, along with other
5 reported biomass materials on the air-dried basis (adb).
6 Table 3. Pyrolysis performance indices of PR and CF with four heating rates program.
7 Table 4. The activation energy (Ea) estimates, according to Model-free Isoconversional
8 methods (KAS, OFW, STR, LTA, FRM, and VYZ).
9 Table 5. Results of Reaction mechanism by Model-fitting Isoconversioanl method (Coats-
10 Redfern method).
11 Table 6. Results of Thermodynamic parameters of PR and CF.

of
12

ro
13 List of Figures
14 Fig. 1 (a). Thermogravimetric (TG) curves of PR.
15 Fig. 1 (b). Thermogravimetric (TG) curves of CF.
-p
re
16 Fig. 2 (a). Differential thermogravimetric (DTG) curves of PR.
lP

17 Fig. 2 (b). Differential thermogravimetric (DTG) curves of CF.


18 Fig. 3 (a). Variation of activation energy (Ea) as a function of conversion rate (α) by different
model-free isoconversional methods during the reaction of PR.
na

19
20 Fig. 3 (b). Variation of activation energy (Ea) as a function of conversion rate (α) by different
21 model-free isoconversional methods during the reaction of CF.
ur

22 Fig. 4 (a). Comparison between theoretical and experimental master plots under a heating rate
Jo

23 of 10 oC/min based on 21 different kinetic models by using Criado-Master plots method of


24 PR
25 Fig. 4 (b). Comparison between theoretical and experimental master plots under a heating
26 rate of 10 oC/min based on 21 different kinetic models by using Criado-Master plots method
27 of CF

39
1st study concerning the pyrolysis kinetics of putranjiva and amaltas non-edible

oilseeds

Both model-free and model-fitting methods were used

Non-edible oilseeds pyrolysis model is discriminated by the master-plot method

Non-edible oilseeds pyrolysis were well represented by An and Rn mechanism

of
ro
-p
re
lP
na
ur
Jo
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like