You are on page 1of 13

Energy Reports 11 (2024) 4277–4289

Contents lists available at ScienceDirect

Energy Reports
journal homepage: www.elsevier.com/locate/egyr

Research paper

Waste to value addition: Utilization of waste corn cob from corn plant
derived novel green acidic catalyst for effective synthesis of esters
Balaji Panchal a, c, Yuzhuang Sun b, Cunliang Zhao a, c, *, Jinxi Wang c, Kai Bian c,
Qiaojing Zhao a, c, Bangjun Liu a, c, *
a
Key Laboratory of Resource Exploration Research of Hebei,Hebei University of Engineering, Handan, Hebei 056038, China
b
Hebei Collaborative Innovation Center of Coal Exploitation, Hebei University of Engineering, Handan, Hebei 056038, China
c
College of Earth Science and Engineering, Hebei University of Engineering, Handan, Hebei 056038, China

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

Keywords: Sulfonation of waste corn cob biomass (CCB) prepared a multifunctional mesoporous acidic catalyst for ethyl
Biomass–derived materials oleate production. For determining the best catalyst synthesis conditions, CCBs were calcineted at different
Mesoporous activated carbon based catalyst temperatures (150–300 ◦ C), sulfonated with 1, 3–propanesulfonate at different temperatures (60–140 ◦ C), and
Calcination
sulfonation times (0.5–4.5 h). The structural, porosity, morphology and surface properties of the samples were
Sulfonation
Catalyst characterization
thoroughly examined by FT–IR spectroscopy, TG analysis, SEM, N2 adsorption–desorption isotherms, XPS
Esterification spectroscopy, and CHNS analysis. As a result of varying reaction conditions in catalyst optimization, including
8 g of 1, 3–propanesulfonate, 3 g of activated carbon CCB250, at 120 ◦ C, and 4 h of reaction time, 94.32% of the
–SO3H was obtained, and the catalyst surface was highly academically active. According to sulfur analysis of the
CCB250–SO3H catalyst, the 120 ◦ C sulfonated activated CCB250 contained –SO3H groups, with a higher density of
active sites of 9.426 mol/g. By sulfonation of 1, 3–propanesulfonate, the mesoporous CCB250–SO3H catalyst was
obtained with an average pore diameter of 3.1–4.8 nm, a pore volume of 1.39–1.02 cm3/g. The mesoporous
CCB250–SO3H catalyst was utilized in the catalytic activity test to esterification with oleic acid. It exhibited a
conversion rate of 94.32% in the production of ethyl oleate. The optimal conditions for this reaction included a
catalyst loading of 150 mg, a molar ratio of ethanol to oleic acid (12:1 mol/mol), a reaction temperature of 80 ◦ C,
and a reaction time of 2 h. Due to the stable attachment of the –SO3H functional group, the mesoporous
CCB250–SO3H catalyst was successfully reused eight times for esterification cycles.

1. Introduction catalysts for efficient biofuel production (Li et al., 2014). In addition to
being affordable, non–toxic, non–corrosive, and readily available,
The excessive consumption of fossil resources has led to serious en­ agricultural waste biomass can serve as catalysts (Guo et al., 2012;
ergy depletion and environmental problems in recent decades. Now, Mares et al., 2021), sugarcane baggasse (Ezebor et al., 2014; Majamo
researchers worldwide are focusing on renewable resources (Zhang et al., 2023), peanut hulls (Abdullaha et al., 2017), biomass waste
et al., 2019). Biofuels are non–toxic, biodegradable, contain no sulfur or (Nda-Umar et al., 2022), lignocellulosic biomass (Zhu S et al., 2020). In
aromatic compounds, and are environmentally friendly fuels that reduce addition to being more sustainable and feasible, biomass–based solid
global warming (Karmakar et al., 2010; Nair, Sivakumar, 2022). The catalysts can also help eliminate agricultural waste biomass disposal
development of eco–friendly materials and bio–inspired catalysts for issues (Naeema et al., 2021).
chemical transformations has gained considerable attention in recent The effectiveness of several organic, metallic, metal oxide and ionic
years within the emerging concepts of green chemistry (Hara et al., liquid catalysts has been thoroughly investigated (Mumtaz et al., 2022).
2010; Chua et al., 2020). The use of biomass–based materials (bio­ The process of homogeneous bases (sodium hydroxide, potassium hy­
polymers, biochar, carbon) has improved biofuel yields and energy ef­ droxide, sodium methoxide, etc.) catalyzed transesterification converts
ficiency, enabling more commercially viable biofuel technologies (Cao biofuel into fuel in a shorter time, but soap formation and catalyst
et al., 2022), and they have also demonstrated great promise as green separation are a few of the major drawbacks (Mendonça et al., 2019). To

* Corresponding authors at: Key Laboratory of Resource Exploration Research of Hebei,Hebei University of Engineering, Handan, Hebei 056038, China.
E-mail addresses: zhaocunling@hebeu.edu.cn (C. Zhao), bangjuliu@126.com (B. Liu).

https://doi.org/10.1016/j.egyr.2024.03.015
Received 13 January 2024; Received in revised form 25 February 2024; Accepted 7 March 2024
Available online 13 April 2024
2352-4847/© 2024 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

convert oil to biofuel, acid catalyst (trans)esterification requires high 2. Materials and methods
temperatures and high pressures with long reaction times (Nath et al.,
2019). It is necessary to develop enzyme catalysts that are ecologically 2.1. Materials
responsible and comply with environmental regulations. As a result,
downstream processing can be reduced since the products do not Corn cobs were provided by a farm in Ninghe District, Tianjin, China.
contain significant impurities from the catalyst (Bastos et al., 2020). A variety of analytical reagent grade solvents were purchased from
Instead of requiring extensive washing steps, solid carbon–based acid Tianjin Kemiou Chemical Reagent Co., Ltd. including 1,
catalysts can simplify the overall process, reduce chemical waste pro­ 3–propanesulfonate, and methanol (AR grade) and ethanol (AR grade)
duction, and maintain high–quality biofuel production at the same time were purchased from Beijing J&K Scientific Ltd. Otherwise; all chem­
(Yadav et al., 2023). The cost of production is further reduced as the icals were obtained in analytical grade. As well as sulfuric acid, potas­
catalyst can be reused for multiple cycles (Mardhiah et al., 2017). Car­ sium hydroxide and sodium chloride, Beijing chemical works, Beijing, P.
bon activated is the product of processing amorphous carbon so that it R.China, supplied oleic acid, phenolphthalein indicator, potassium
has a large surface area for adsorption and chemical reactions (Hamza phthalate and sodium hydroxide.
et al., 2020).
In this study, corn cob activated carbon (AC) is sulfonated with 1, 2.2. Sample preparation
3–propanesulfonate to prepare an acidic catalyst (CCB250–SO3H) that
may be used to produce ethyl oleate from oleic acid. The majority of the A drying procedure at 100 ◦ C was performed overnight on corn cob
literature studies focus on catalyst preparation using calcination–sulfo­ biomass samples. After drying, the samples were grand by grander to the
nation methodologies without examining how reaction conditions affect final particle size under 200 µm. Fine particle samples of corn cobs were
–SO3H density. A two–step method has several disadvantages, including analyzed for moisture–by–moisture analyzers (KERN DBS, Germany).
being a complicated and costly process and resulting in the loss of An aluminum plate was weighed with approximately 30 g of a fine
adsorption capacity. In contrast, a one–step method can maintain a high particle corn cob sample, and the sample was heated to constant weight
specific surface area and stable properties of the raw material at 105 ◦ C under a vacuum oven. Afterwards, the dried sample was used
throughout the process of preparing activated carbon (Zhang et al., to prepare the catalyst.
2016; Fu et al., 2021). Oginni, Singh, 2019 explained that the synthesis
of activated carbon via a chemical activation pathway can be divided 2.3. Catalyst synthesis procedure
into one–step or two–step activation processes. In this work, a one–step
method was conducted. In addition to cellulose (C6H10O5), hemicellu­ Step I: The raw CCB was chemically activated with 30% concentrated
lose (C5H10O5), and lignin (C11H14O4, 14–26%), corn cob contains a hydrochloric acid (HCl) in this study because it facilitates high porosity
high amount of carbon, as well as high thermal and biodegradability and specific surface area development at low temperatures (150–300
properties (Lu et al., 2017), as well as some amounts of ash (Poveda-­ ◦
C). A typical process involves soaking 25 g of raw CCB in 30% conc HCl
Giraldo et al., 2021). As a low–cost and abundant renewable energy (raw CCB: 30% conc HCl, 1:2 w/w ratio) impregnation ratio for 24 h.
resource, this is the most common and most abundant in the world Following that, the material is transferred to a quartz crucible and
(Miranda et al., 2018;Yuan et al., 2021). A variety of solid acids based on calcined at 150–300 ◦ C under nitrogen flow for 2 h. The raw CCB
sulfonated carbon have been extensively investigated (Zeng et al., calcination was carried out at the rate of 10 ◦ C min–1 until the desired
2013). It is well known that solid acids can be easily operated, have good temperature was achieved. A rate of 10 ◦ C min–1 was applied to the CCB
catalytic efficiency, and are highly selective in catalysis, which reduces calcination until the desired temperature was achieved. Activated car­
the cost of biofuel production (Dhawane et al., 2016a,b). In a one–step bon (CCB) is generated by washing the black solid produced by the
method involving simultaneous calcination and sulfonation, Guo et al. process with hot double–distilled water until pH 7 and then drying it at
(2012) prepared a similar sulfonated carbon catalyst from lignin derived 105 ◦ C to yield CCB. A weighted average of yields was determined at
from nutshells of Xanthoceras sorbifolia (an oil plant in China). Hara et al. 20%, 26%, 35%, and 33%, respectively, at 150, 200, 250, and 300 ◦ C.
(2010) developed a method of preparing it involving sulfonation of There are four different types of activated carbon (CCB), CCB150,
incompletely pyrolyzed carbonaceous materials obtained from biochar CCB200, CCB250, and CCB300, where CCB represents numbers denoting
(Yu et al., 2010), wood (Kitano et al., 2009), and canola meal (Rao et al., the calcination temperatures.
2011). Step II: The activated carbon materials were then powdered using a
A sulfonated carbon–based catalyst (CCB250–SO3H) synthesized with mortar and pestle. In this study, activated carbon was combined with
1, 3–propanesulfonate is investigated in the present study under Folch mixture (1:2 v/v methanol: chloroform) 1:3 w/v for 12 h and
different experimental conditions, including activated carbon CCB250, 1, subsequently filtered with Whatman filter paper. Separate the liquid
3–propanesulfonate volume, sulfonation temperature, and sulfonation phase from the precipitated product. A vacuum oven was then used to
time. Material with mesopore structure exhibits high surface area, nar­ dry the precipitated product.
row pore size distribution, and large pore volume, allowing high con­ Step III: Sulfonation is achieved by mixing the activated carbon
centrations of –SO3H groups and good access to active sites for CCB250 product (1 g to 4 g) with 1, 3–propanesulfonate (2 g to 10 g). The
sulfonation (Shuit et al., 2013). Additionally, the catalyst surface is mixture was mixed and heated to the required temperature (60–140 ◦ C)
hydrophobic, which means it can adsorb long–chain organic molecules at the various set times (0.5–4.5 h). In order to remove the entire
such as free fatty acids without being deactivated by water adsorption. remaining sulfonic acid agent from the catalyst, ethanol was added
As part of the physicochemical characterization of the CCB250–SO3H following each cycle. A vacuum oven was used to dry out the final
catalyst, Fourier transform infrared (FT–IR), scanning electron micro­ catalyst (CCB250–SO3H) overnight at 60 ◦ C. Esterification and acid
scope (SEM), thermogravimetric analysis (TGA), N2–adsorption and density were used to confirm catalyst activity.
desorption isotherms, X–ray photoelectron spectroscopy (XPS), car­
bon–hydrogen nitrogen sulfur (CHNS), and phenolphthalein indicator 2.4. Catalyst characterization
analysis was conducted. The CCB250–SO3H catalyst was further evalu­
ated to determine whether it could be used without further treatment for FT–IR analysis of raw CCB, calcined CCB250, and CCB250–SO3H was
esterification reactions. performed using KBr pellets using a Nicolet 510 P spectrometer
(Thermo, AVATAR 370, USA). A Tescan VEGA 3 LMU microscope was
used to examine the catalyst morphology using SEM. Based on the linear
portion of the Brunauer Emmett–Teller BET, the specific surface area

4278
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

was calculated, and the total pore volume was measured. A Thermo 3. Result and discussion
KAlpha X–ray photoelectron spectrometer was used for the XPS ana­
lyses. Thermal decomposition patterns of the raw CCB, calcined CCB250 3.1. Effect of calcination temperature on the structure of CCB
and sulfonated CCB250–SO3H catalyst in an inert atmosphere were
evaluated using TGA–5500 thermogravimetric analyzer. An X–ray During the catalyst preparation process, raw CCB was pretreated
diffractometer (BRUKER–D8) was used to analyze raw CCB, calcinated with 30% HCL acid (Table 1). A comparison was made between the raw
CCB250 and sulfonated CCB250–SO3H catalyst using scanning step length CCB, CCB150, CCB200, CCB250, and CCB300 and the increment in surface
and scanning speed of 0.03◦ and 0.2 s/step, respectively (Stefanidis composition and structure. According to the analysis results, the calci­
et al., 2016). Loss on ignition (LOI) was determined by igniting 2 g nation temperature of 250 ◦ C has a great impact on the pore structure of
samples in a 1.5 kW laboratory furnace for 1 h at 1000 ◦ C. Raw and activated carbon. Volumes of the pores 0.07 cm3/g, 1.02 cm3/g,
calcined CCB250, and the final produced catalyst (CCB250–SO3H) were 1.15 cm3/g, 1.39 cm3/g, and 1.31 cm3/g for raw CCB, CCB150, CCB200,
analyzed by an elemental analyzer from CHNS (CKIC 5E-CHN2200). CCB250, and CCB300, respectively. As compared to raw CCB, CCB150,
CCB200, CCB250, and CCB300, CCB250 showed a larger surface area of
2.5. Acid density analysis 398 m2 g–1. In addition, the diameter of the pores increased from
1.85 nm for the raw CCB to 3.72 nm for the calcinated CCB250. The raw
A sulfonated catalyst was determined to have a strong acid density CCB has clearly improved as a result of the calcination process. One
(AD, mmol/g) of –SO3H. In a conical flask, 250 mg of CCB250–SO3H advantage of 250 ◦ C is that it increases surface area, porosity, and
catalyst was mixed with 30 mL of 0.5 M NaCl solution. Magnetic stirring diameter, thus providing more sulfonation sites. Activated carbon,
was used to exchange H+ in the form of –SO3H for Na+ for 2 h, followed however, becomes harder with higher calcination temperatures (300
by filtering. The filtrate was titrated using 0.1001 mol/L NaOH standard ◦
C), resulting in a drastic reduction in pore volume, leading to fewer
solutions with phenolphthalein as an indicator. It took 30 seconds for active sites for grafting of sulfur radicals, resulting in lower acid density.
the colorless filtrate to turn pink, and to remain at this color for at least The surface morphology of calcined samples is affected by the
30 seconds. At the same time, a blank test was performed. As a result, the calcination temperature, according to Boro et al. (2012). According to
following formula was used to calculate the accurate strong acid density: Kim and co–workers, xylan (hemicellulose), cellulose and lignin
decompose at temperatures above 150 ◦ C, 220 ◦ C, and 250 ◦ C during
CNaOH × (V1 − V0 )
AD− SO3 H = (1) carbonization; xylan (hemicellulose) hardly decomposes at all (Volpe
mCatalyst
et al., 2020). According to Betiku et al. (2017), cocoa pod husk was
calcined and the resultant surface area was 2.76 m2 g–1, the pore volume
where AD− SO3 H represents the strong acid (–SO3H) density of the
was 0.012 cm3/g, and the pore size was 1.77 nm. Activated carbon (AC)
catalyst (mmol/g); CNaOH represents the concentration of the NaOH
is an ideal support for acidic molecules as it has a network of micro,
standard solution; V1 and V0 are the quantities of NaOH standard so­
meso, and macropores, all interconnected, making it ideal for embed­
lution consumed during the catalyst and blank tests, respectively; and
ding acids (Dhawane et al., 2016a,b). Similar observations have been
mcatalyst represents the mass of catalyst.
reported for coconut husks (Vadery et al., 2014), and bagasse–derived
palm trunks (Ezebor et al., 2014). Using Acacia nilotica wood calcined
2.6. Catalytic testing by esterification
for 3 h at different temperatures, Sharma et al. (2012) synthesized a
heterogeneous catalyst with a surface area ranging from 1.3 to 3.7 m2
The esterification of synthesized sulfonated catalyst determines their
g–1. Consequently, 250 ◦ C is selected as the correct calcination tem­
activity. Using oleic acid and ethanol as a model, each synthesized sul­
perature. The stability of catalysts is not significantly enhanced when
fonated acid catalyst was tested for its ability to form ethyl oleate. An
calcination duration is further extended, as shown by (Lou et al., 2012)
ethanol–cooled condenser, a magnetic stirrer, and a digital thermometer
within a range of calcination duration at 0.5–20 h.
were used to conduct the esterification reactions in a 100 mL conical
glass flask. A reaction with oleic acid was conducted with synthesized
3.2. Optimization of sulfonation factors
sulfonated catalyst (150 mg) at their ethanol reflux temperature (about
80 ◦ C) for 2 h using 2.82 g of oleic acid and 16.56 g of ethanol. Once the
3.2.1. Effect of activated carbon CCB250 amount on sulfonation
reaction was complete, the glass flask was removed from the bath and
Catalyst synthesis requires a significant amount of activated carbon
allowed to cool to room temperature. A centrifugal separator was used
CCB250. The activated carbon CCB250 was varied from 1 g to 4 g, while
to separate the products from the catalyst. The ethyl oleate phase was
then heated in a drying oven after being separated from the aqueous and
organic phases. Ethyl oleate acidity was determined by titration with Table 1
Properties of calcinated CCB with 150 ℃ to 300 ℃ temperature.
0.01 mol/L potassium hydroxide (in triplicates). As a result of the con­
version, the following result was obtained: Samples SBET Pore Acid density References
(m2g− 1) volume (mmol/g− 1)
(Ci –Ct ) (cm3/g)
Cest (%) = (2)
Ci Raw CCB 89 0.07 0.03 This work
CCB150 310 1.02 1.09 This work
where Ci is the initial acid value and Ct is the acid value measured at CCB200 372 1.15 1.22 This work
time t. The acid value was calculated as follows: CCB250 398 1.39 1.42 This work
CCB300 380 1.31 1.4 This work
56.1 × N × V WVB–20–SO3 113 0.63 N.D Kastner et al.,
Acid value(mg KOHg–1 ) = (3) 2012
W
SO3H–CSR–400–180 53 0.007 2.12 Wang et al.,
where N = strength of KOH (in normality); V = volume of KOH 2014
Activated 1131 N.D N.D Naeema
consumed; W= weight of sample.
carbon–SO3H et al., 2021
OMC–130–SO3H 813 0.69 1.83 Liu et al.,
2008a,b
Corncob–SO3H 80 N.D 0.16 Arancon
et al., 2011

N. D–Not detected

4279
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

the 1, 3–propanulfonate was used at a sulfonation temperature of 120 ◦ C catalytic performance, and the esterification yield (%) remains around
and a sulfonation time of 4 h (Fig. 1). The high yield of 94.32% was 94.32%. Tao et al. (2015) prepared sulfonated carbon catalyst from
obtained in 3 g of activated carbon CCB250, further, expansion did not waste biomass by carbonizing bagasse at 648 K for 0.5 h, sulfonating it
enhance the acid density or yield (Table 2). In contrast, 4 g of activated at 423 K for 15 h, and receiving an –SO3H density of 1.06 mmol/g.
carbon CCB250 produced the same yield (%). A yield of 41.25% was Catalytic yield plays an important role in evaluating the effectiveness of
recorded at 1 g. In electrophilic aromatic substitutions, sulfonation in­ the sulfonation process, but it is more important to incorporate sulfur
volves replacing hydrogen atoms in aromatic hydrocarbons with sul­ sites into the CCB250–SO3H catalyst since catalytic activity is attributed
fonic acid functional groups. A growing interest in carbon–based to the sulfonic acid (–SO3H) functional groups (Zhang et al., 2015). As a
materials is being driven by the ease of sulfonation of biomass and result of sulfonation for 4 h, fewer –SO3H active sites are formed,
related wastes (Lu et al., 2016). Therefore, we considered 3 g to be an resulting in a more rigid structure.
optimal amount of activated carbon CCB250 for sulfonation. According
to Akinfalabi et al. (2017), palm seed cake was sulfonated for 12 h at 150 3.2.4. Effect of sulfonation temperature

C, and 91% of the fatty acids methyl ester was converted. Sulfonation of activated carbon CCB250 is conducted at various
temperatures between 60 ◦ C and 140 ◦ C and the catalyzed esterification
3.2.2. Effect of 1, 3–propansulfonate volume efficiency is shown in Fig. 4. A higher sulfonation temperature results in
A slight increase in esterification yield occurs when 1, more –SO3H groups loading and a higher esterification yield. Activated
3–propanesulfonate is sulfonated with 8 g (Fig. 2). Additionally, 1, carbon CCB250 amount (3 g), sulfonation time (4 h), 1,
3–propanesulfonate volumes over 8 g reduce acid density. In addition to 3–propanesulfonate volume (8 g), and agitation speed of 100 rpm were
graphitizing biomass and cross linking carbon sheets, increasing the 1, kept constant. A yield of 94.32% was achieved when the sulfonation
3–propanesulfonate yields a more rigid carbon framework and a high temperature was 120 ◦ C, contributing to high surface area and pore
density of sulfonic acids, resulting in the elimination of carbon­ volume. When the sulfonation temperature is raised above 120 ◦ C, the
–hydrogen bonds prior to the attachment of the sulfonic acid groups. In yield and acid density are significantly reduced (Table 2). At higher
(Table 2), it can be seen that the acid density of the CCB250–SO3H sulfonation temperatures, CCB250–SO3H catalyst becomes harder,
catalyst increased with increasing propanesulfonate volume, with 8 g of resulting in fewer sites for radical grafts (Kang et al., 2013). At lower
propanesulfonate volume achieving the highest yield (94.32%) and sulfonation temperatures, there is no obvious change in surface area or
–SO3H group density (94.32%). In contrast, the acid density for pore volume for CCB250–SO3H catalyst, suggesting that the framework
CCB250–SO3H catalyst is clearly altered when 8 g of 1, 3– propane­ of catalyst remains intact. When sulfonating aromatic compounds, the
sulfonate is added, suggesting that the catalyst’s framework remains sulfonic groups location is influenced by the sulfonation temperature in
intact. After adding 8 g of 1, 3–propanesulfonate volume, no significant a reversible electrophilic substitution reaction (Yu et al., 2016). Savaliya
change in catalytic performance occurs, and the esterification yield is and Dholakiya (2015) report a 25–40% yield by synthesizing SAC–SCB
94.50%. For the esterification of oleic acid, Liu et al. (2013) prepared at 180 ◦ C for 10 h. At 120 ◦ C, the best yield was achieved. A high
sulfonated corn straw–activated carbon. A number of previous studies temperature can result in the collapse of the carbon skeleton and the
have found that sulfonated activated carbon material is the most effi­ occurrence of side reactions, thereby reducing the introduction of –SO3H
cient and economical catalyst for the synthesis of esters. According to groups. At a lower temperature, however, the sulfonated product is
Konwar et al. (2014), this catalyst possesses indistinguishable acidic unstable and easily decomposed (Yu et al., 2017). Due to this rigid
properties. carbon framework, sulfonic acid is highly concentrated. The sulfur
content of CCB250–SO3H catalyst ranged from 0.19% to 8.74%, indi­
3.2.3. Effect of sulfonation time cating successful sulfonation of –SO3H into aromatic carbon rings (Liu
In Fig. 3, a 0.5 h to 4.5 h sulfonation time significantly improved the et al., 2013).
catalytic capability of the CCB250–SO3H catalyst, while the other vari­
ables were fixed at 1, 3–propansulfoante volume (8 g), activated carbon 3.3. Characterization of CCB250–SO3H catalyst
CCB250 (3 g), and 120 ◦ C for sulfonation. The yield (%) increased from
60% to 94.32% as the sulfonation time was increased from 0.5 to 4 h 3.3.1. Acid density of synthesized catalyst
(Table 2). After 4 h of sulfonation time, there is no significant change in As one of the important parameters of acid catalysts, acid density
directly reflects the number of catalytic sites. The acid density of
CCB250–SO3H catalyst was prepared from different conditions such as
the amount of activated CCB250 carbon, sulfonation time, sulfonation
temperature, and volume of 1, 3–propaneulfonate (Table 2). Sulfonated
CCB250–SO3H catalyst showed an increase in acid density from
0.27 mmol/g to 9.426 mmol/g, resulting in a reduction of specific sur­
face area because –SO3H groups are attached to the activated carbon
surface. CCB250–SO3H catalyst acidified via calcination appears to have
the highest accessibility of sulfonated moieties thinks to its strong
dependence on the calcination method. In spite of the low surface area of
the catalyst, its total acid density was 9.426 mmol/g. Due to the large
pores and large pore volume of the catalyst, 1, 3–propanesulfonate was
easily incorporated into the carbon bulk of the catalyst, creating cova­
lent bonds between –SO3H groups and the surface area. Fraile et al.
(2014) calculated the acidity due to sulfur groups, when all sulfur was
present in the oxidation state, assuming sulfur was all present in the
oxidation state. The density of –SO3H is a key factor that determines the
catalytic activity of the catalyst (Csac) in esterification reactions of FFA
Fig.1. Effect of activated carbon CCB250 amount in the synthesis of the (Malins et al., 2016). As more acid sites are present in the catalyst, a
CCB250–SO3H catalyst: Sulfonation temperature 120 ◦ C, sulfonation time 4 h, higher acid capacity is expected, resulting in better catalytic ability in
1,3–propaneulfonate volume 8 g and various amounts of activated carbon the production of the ester process (Mardhiah et al., 2017). Due to the
CCB250 of 1 g to 4 g. high –SO3H density, the sulfonation method was found to be more

4280
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Table 2
Acid density of CCB250–SO3H catalyst prepared in different conditions.
Sr. no. CCB250 amount 1, 3–propanesulfonate (g) Reaction time Reaction temperature Acid density (mmol/ References
(g) (h) (℃) g)

1 3 2 4 120 4.025 This work


2 3 4 4 120 6.267
3 3 6 4 120 7.241
4 3 8 4 120 9.426
5 3 10 4 120 9.368
6 4 8 4 120 9.393
7 2 8 4 120 3.209
8 1 8 4 120 2.873
9 3 8 4.5 120 9.318
10 3 8 3.5 120 8.512
11 3 8 3 120 7.863
12 3 8 2.5 120 5.772
13 3 8 2 120 3.984
14 3 8 1.5 120 3.084
15 3 8 1 120 2.153
16 3 8 0.5 120 1.861
17 3 8 4 140 9.243
18 3 8 4 100 7.906
19 3 8 4 80 5.231
20 3 8 4 60 3.762
SCB–SO3H 2 200 mL 5 300.00 5.63 Akinfalabi et al.,
2017
Sulfonated activated 6 10 g 6 60.00 89 Han et al., 2022
carbon

Sugarcane bagasse biomass (SCB–SO3H)

Fig. 2. Effect of 1, 3–propanesulfonate in the synthesis of CCB250–SO3H cata­ Fig. 3. The effect of sulfonation time on CCB250–SO3H catalyst synthesis: Sul­
lyst: Sulfonation temperature (120 ◦ C), sulfonation time (4 h), activated carbon fonation temperature (120 ◦ C), activated carbon CCB250 (3 g), 1, 3
CCB250 (3 g), and various amount of 1,3– propanesulfonate of 2 g to 10 g. –propanesulfonate (8 g), and various sulfonation times of 0.5 h to 4.5 h.

efficient in this study (Arancon et al., 2011; Prabhavathi Devi et al., modes of CO–2 3 (Han et al., 2022). Raw CCB also showed distinct and
2014). It has been reported in Kastner et al. (2012) that sulfonated minor sharp peaks at 3416.9, 2923.44, 2854.26, 1735.82, 1635.18,
wood–based activated carbon catalysts with lower –SO3H acid sites 1515.18, and 1382.69 cm− 1. According to Fig. 5b, characteristic IR
density (0.2 –0.01 mmol/g) had greater esterification activity than a peaks of activated carbon CCB250 correlate with vibrations of asym­
sulfonated peanut hull char catalyst with a higher –SO3H acid site metric C–O bonds around 1263.30, 1162.78, 1108.14 and
density (0.61–0.03 mmol/g). 1056.97 cm–1. Furthermore, hydroxyl (OH–1) groups correlate with
557.87 cm–1 and can be inscribed to the bending mode of the (OH–1)
3.3.2. Surface functionality group. After calcination, the reduced mass of the functional group
In Fig. 5, FT–IR patterns have been obtained for raw CCB, activated attached to the CO–23 ions decreases while the intensity of CaCO3 peaks
carbon CCB250, and sulfonated activated carbon CCB250 (CCB250–SO3H). declines. The major absorption bands of raw CCB existing at 1382.69,
There is a broad peak at 3416.9 cm–1 in the spectrum, which corre­ 1250.85, 1161.87, and 607.15 cm− 1 almost disappeared in calcinated
sponds to the –OH group, which is due to adsorbed water molecules on CCB250 which clearly indicates the effectiveness of selected calcination
raw CCB, activated carbon CCB250, and CCB250–SO3H catalyst surfaces temperature which degraded the complex carbohydrate–lignin matrix.
(Park et al., 2008; Etim et al., 2018). In Fig. 5a, the absorption peaks of Sulfonated CCB250–SO3H catalyst has significantly different spectra
the raw CCB peaks occurred at 1250.85 cm–1, 1161.87 cm–1, from activated carbon CCB250, and raw CCB. The peaks at 1350.12,
1049.21 cm–1, and 607.15 cm–1 which might indicate that raw CCB 1125.6, 1025.56, 897.64, 790.68, 710.45, and 530.40 cm–1 only appear
biomass contains carboxylate cellulose, hemicellulose and lignin, as well in the spectra of sulfonated CCB250–SO3H catalyst (Fig. 5c). A successful
as C–O asymmetric stretching and out–of–plane bending vibration sulfonation process and the presence of –SO3H group on the surface of

4281
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

the FT–IR spectra, it was found that grafting –SO3H groups onto
CCB250–SO3H catalyst using the sulfonation method was feasible (Zeng
et al., 2014).

3.3.3. Morphologies of CCB250–SO3H catalyst


By scanning electron microscope (SEM), the surface morphology of
raw CCB, activated carbon CCB250, and CCB250–SO3H catalyst was
examined (Fig. 6). Clearly, in Fig. (6a), the rough surface is due to the
carbohydrates. Fig. (6b), shows that the activated carbon CCB250 has a
large number of mesopores on its surface. Activated carbon CCB250
exhibited irregular surfaces with well–developed porous structures.
Pores are created as a result of super heating, which increases the surface
area and irregularities as a result of sulfonation (Table 3). These pores
allowed for a higher volume of 1, 3–propanesulfonate to enter the car­
bon bulk, which resulted in a higher covalent bond between the carbon
and the –SO3H group (Inagaki, 2009). As shown in Fig. 6c, CCB250–SO3H
Fig.4. Effect of sulfonation temperature in the synthesis of CCB250–SO3H catalyst, after sulfonation, have an irregular surface with well­
catalyst: Activated carbon CCB250 (3 g), 1, 3–propanesulfonate (8 g), sulfona­ –developed porous structures with average diameters of 3.1–4.8 nm,
tion time (4 h), and various sulfonation temperature of 60 ◦ C to 140 ◦ C. exhibiting impermeable porous surfaces. SEM images revealed aggre­
gates with the mesoporous structure of the catalyst, which indicates the
sulfonated CCB carbons were confirmed by the emergence of additional presence of CCB250–SO3H catalyst during the synthesis process, even if
bands at 1125.6 cm− 1 and 1025.56 cm− 1 (S– –O stretching), 897.64, the arrangement is unchanged. SEM observations indicate that the
790.68, 710.45, and 530.40 cm− 1 (–SO3H stretching) on CCB250–SO3H calcination and sulfonation processes significantly change the catalyst’s
catalyst IR spectra. According to the vibration bands, the –SO3H func­ morphology. In the sulfonation process, with a high density of functional
tional group was impregnated into the carbon surface (Ramadoss, groups such as –SO3H, COOH, and phenolic –OH groups (Ngaosuwan
Muthukumar, 2016; Li et al., 2012). Bora et al. (2018) report that the et al., 2016), a change in the chemical nature of the backbone was ob­
FT–IR band at 1120 cm–1 indicates the presence of the O– –S––O group in tained with a decrease in aliphatic chains, an increase in aromatic rings,
the structure. In CCB250–SO3H catalyst, the major absorption bands of and an increase in oxidized groups.
activated carbon CCB250 were almost absent, indicating the sulfonation
of –SO3H groups. Moreover, the peak at 530.40 cm–1 indicates the 3.3.4. SBET and pore volume of CCB250–SO3H catalyst
presence of S–O groups on the surface of functionalized CCB250–SO3H Sulfonation temperature, specific surface area, pore volume, and
catalyst, while the peak at 1025.56–710.45 cm–1 was attributed to average pore diameter are complexly correlated, according to previous
asymmetric and symmetric O– –S– –O stretching vibrations. The low studies. This was achieved by analyzing raw CCB, activated carbon
intense band at 530.40 cm− 1 signifies the presence of the sulfur group in CCB250, and CCB250–SO3H catalyst using BET. The CCB250–SO3H cata­
CCB250–SO3H catalyst which is in good agreement with the sulfonation lyst show type–IV isotherms for N2 adsorption and desorption (Fig. 7a
results. The functional groups present in CCB250–SO3H catalyst indicate and b), characteristic of mesoporous materials (Zhao, Collinson, 2010).
that it can be used for esterification for ester production. As a result of The presence of hysteresis loops, which were nitrogen condensation

Fig. 5. FT–IR spectrum of raw CCB (a), activated CCB250 (b), and CCB250–SO3H catalyst (c).

4282
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Fig. 6. SEM images of raw CCB (a), activated carbon CCB250 (b), and sulfonated CCB250–SO3H catalyst (c).

Table 3
Surface properties of mesoporous CCB250–SO3H catalyst.
Catalysts Total SO3H Specific Pore References
acid density surface volume
density (mmol/ area (cm3/g)
(mmol/ g− 1) (b) (m2g− 1)
g− 1) (a)

Raw CCB 0.46 0.19 89 0.07 This work


Activated 0.27 0.09 398 1.39 This work
carbon
CCB250
CCB250–SO3H 9.426 8.74 306 1.02 This work
CCB250–SO3H 7.404 5.1 150 0.5 This work
(cycled)
Starch–SO3H 4.13 1.5 N.D N.D Zong et al.,
2007
Corncob–SO3H N. D 0.16 80 N.D Arancon
et al., 2011
ACPhSO3H N.D 0.72 114 0.27 Malins
et al., 2016
IC–SSCB 0.9 52.13 0.054 Ezebor
et al., 2014

where, (a) based on titration. (b) based on elemental analysis.


N. D-not detected

steps, was observed at high P/Po around 0.0–1.0, indicating the pres­
ence of multi modal pores. Aggregation synchronized with large particle
size led to low surface area of the catalysts (Liu et al., 2008a,b). A larger
particle size at high calcination temperature results in a reduced surface
area of the catalyst, which can be verified by the BET surface area re­
sults. As a result, the surface area of activated carbon CCB250 was
significantly increased from 89 m2 g–1 to 398 m2 g–1 and the volume of
pores was significantly diminished from 1.02 cm3/g to 1.39 cm3/g,
Fig.7. N2 adsorption and desorption isotherms of sulfonated CCB250–SO3H
respectively, when activated carbon CCB250 was sulfonated by 1,
catalyst (a) and pore size distribution of CCB250–SO3H catalyst (b).
3–propanesulfonate, the SBET and the pore volume of the samples greatly
decreased (from 398 m2 g–1 to 306 m2 g–1 and from 1.39 cm3 /g to
from the destruction and closure of numerous micropores on the surface
1.02 cm3 /g, respectively). The mesoporous CCB250–SO3H catalyst was
of activated carbon CCB250 during sulfonation.
obtained with an average pore diameter of 3.1–4.8 nm. This may result
In addition to its high surface area, the CCB250–SO3H catalyst has a

4283
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

porous structure with an average pore size of 4.8 nm, indicating the
presence of some mesopores that are conducive to macromolecule
diffusion such as ethyl oleate and oleic acid (Waqas et al., 2018),
resulting in high activity. After sulfonation, BET area, pore size, and pore
volume of CCB250–SO3H catalyst decreased dramatically compared to
raw CCB to activated carbon CCB250 (Fig.7, Table 3). The larger surface
found in the CCB250–SO3H catalyst prepared to indicate a linear rela­
tionship between the surface areas and the catalytic activity (Savaliya
et al., 2015; Gohain et al., 2017), this CCB250–SO3H catalyst can be an
effective alternative to other conventional catalysts because it has a
better interaction between catalyst active sites and reactants.

3.3.5. Elemental analysis (CHNS analysis) of sulfonated CCB250–SO3H


catalyst
The elemental analysis (Table 4) indicates that the sulfur content in
activated carbon materials is 0.09% before and 8.74% after sulfonation,
indicating that –SO3H is successfully introduced into the aromatic car­ Fig. 8. TG analysis pattern of mesoporous CCB250–SO3H catalyst.
bon rings. Sulfonated CCB250–SO3H catalyst was found to have func­
tional groups on the surface based on elemental analysis. It appears that the thermal decomposition of COOH, OH and –SO3H groups of Csac
grafting the –SO3H to replace the C–H bonds in the raw CCB resulted in a obtained from lignocellulosic material occurs at temperatures above
higher sulfur content, while a lower C, H, and N content (Lin et al., ~150 ◦ C.
2019). In addition to that, the carbon (51.39%), oxygen (35.1%), ni­
trogen (0.35%), and sulfur (8.74%) analyses of the CCB250–SO3H cata­ 3.3.7. CCB250–SO3H catalyst by XPS analysis
lyst prepared in this study are significantly higher than those obtained The chemical structures of the mesoporous CCB250–SO3H catalyst
from the Bamboo–SO3H catalyst (such as carbon = 78.11%, oxygen = was detected by XPS analysis in Fig. 9. Peaks of C1s, O1s, N1s, and S2p
19.69% and sulfur = 2.20%) prepared by (Farabia et al., 2019). were obviously found in the XPS of the mesoporous CCB250–SO3H
catalyst, as shown in Fig. 9a, which is consistent with the elemental
3.3.6. Thermal stability of CCB250–SO3H catalyst composition of the mesoporous CCB250–SO3H catalyst. In the C1s
A thermal and oxidative stability test was performed on the catalyst spectrum of mesoporous CCB250–SO3H catalyst, it could be seen that
using the TGA. Above 120 ◦ C, the catalyst degrades significantly. As comprises of two peaks with different binding energies via deconvula­
shown in Fig. 8, the curve represents the percentage weight loss of tion. The C1s spectrum was split into two peaks at 284.69 eV and
CCB250–SO3H catalyst as temperature increases. Three distinct weight 285.88 eV, associated with the C–C and C–N groups, respectively. This
loss regions can be seen in the TGA pattern of the sulfonated CCB250. indicates the presence of neutral nitrogen and positively charged ni­
Water and free moisture adsorbing on the surface of the CCB250–SO3H trogen (Fig. 9b). The O1s signal appears as a large and asymmetrical
catalyst was the cause of the first, more rapid loss between 50 and 160 peak centered at 529.4 eV. It can be deconvoluted into two components

C. Another weight loss region occurred between 160 and 550 ◦ C; 25.2% at 513.57 eV and 533.11 eV, which are assigned to the C–O or C– –O
of the weight loss in this region was attributed to the thermal decom­ bonds, respectively (Fig.9c). The N1s peak was fitted with a component
position of oxygen–containing groups from catalyst frameworks. During at 401.56 eV, which corresponded to oxides and carbonates, respec­
the temperature range of 200–300 ◦ C, the catalyst was relatively stable, tively ((Fig. 9d). Fig. 9e, shows the (S2p) XPS spectrum of the
with only a slight decrease in weight of 4.5%. A rapid loss of residue CCB250–SO3H catalyst. The peak at 70.95 eV is attributed to sulfur in
carbon deposition was observed at 350–700 ◦ C, resulting in a rapid –SO3H groups. According to these results, CCB250–SO3H catalyst con­
decrease of weight. Compared to the previously reported sulfonated tains the –SH, S–
–O, and –SO3H groups, with the majority of the sulfur
carbon catalyst (Malins et al., 2016), the reported catalyst show excel­ corresponding to sulfonic acid groups (>80%). As shown in Fig. 9, as
lent thermal stability under nitrogen conditions. Overall, CCB250–SO3H seen in the spectra, carbon (51.39%), oxygen (35.1%), nitrogen
catalyst demonstrated good performance at temperatures near 250 ◦ C (0.35%), and sulfur (8.74%) are present near the surface. As shown in
and exhibited good operational stability. According to TGA results, the the Table 5, CCB250–SO3H catalyst has been analyzed in terms of its
catalyst prepared from CCB250–SO3H catalyst exhibits better stability elemental composition. Sulfonation results in the carbon materials
due to its more ordered structure (higher aromaticity) and high surface containing –SO3H groups, and they contain all sulfur atoms in –SO3H
functional group concentration. Santos et al., (2015) also observed that groups after the sulfonation process. XPS analysis was conducted on
CCB250–SO3H catalyst to determine its surface chemical composition.
The valence states and peak values of C1s and S2p of the catalyst also
Table 4
Elemental analysis of raw CCB, activated carbom CCB250 and CCB250–SO3H promote a high level of –COOH generation during catalyst preparation
materials. (Tang, Niu, 2019). As a result, Cagnon et al. (2009) found that carbon
precursors with lower carbon content and higher oxygen content tend to
Catalysts C(%) H N O(%) S References
(%) (%) (%)
have a bigger surface area than carbon precursors with a higher carbon
content and lower oxygen content. According to FT–IR data, the catalyst
Raw CCB 47.71 5.5 0.51 46.1 0.17 This work
surface contains sulfonic, carboxylic, and phenolic groups (Zeng et al.,
activated carbon 51.1 5.16 0.59 43.05 0.09 This work
CCB250 2016).
CCB250–SO3H 51.39 4.51 0.35 35.1 8.74 This work
CCB250–SO3H 39.8 3.35 0.08 26.7 8.75 This work 3.4. Catalytic performance of CCB250–SO3H catalyst on various free fatty
(cycled)
MAC–SO3H 54.65 3.73 4.49 34.76 2.35 Konwar et al.,
acids
2013
Solid carbon–based 41.4 4.46 N.D. 29.7 7.81 Liu et al., It was demonstrated that the CCB250–SO3H catalyst produced oleic
sulfonic acid 2013 acid, linoleic acid, linolenic acid, and lauric acid when esterified with
PAC–600 65.6 2.21 0.72 28.85 2.62 Liu, 2020
ethanol. An examination of reaction conditions was conducted by
N.D–Not detected loading a CCB250–SO3H catalyst on the bases of free fatty acid, FFA

4284
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Fig. 9. Image shows an XPS survey scan (a), a high–resolution C1s XPS spectrum (b), a high–resolution O1s XPS spectrum (c), a high–resolution N1s XPS spectrum
(d), and a high–resolution S2p XPS spectrum (e) of the CCB250–SO3H catalyst.

myristic, palmitic, stearic, linoleic, and oleic acids were obtained under
Table 5
reaction conditions (Table 6). In addition to its strong acidity and a
Surface element concentration (atom %) from XPS.
greater number of acid sites, CCB250–SO3H catalyst exhibits remarkable
Products O1s C1s N1s S2p catalytic properties. As shown in (Table 7), the maximum esters yield
Raw CCB 24.46 70.83 3.63 0.08 obtained for the CCB250–SO3H catalyst was actually higher than that of
Activated carbon CCB250 20 69.38 10.37 0.25 most other popular solid acid catalysts when compared to those reported
CCB250–SO3H 35.1 51.39 0.35 8.74
in the literature for other sulfonated biochar catalysts. As a result of their
high density of –SO3H, mesoporous carbon–based catalysts have a high
(mmol), and ethanol (mmol), and allowing the reaction to proceed for ability to intercalate hydrophilic molecules (Hara et al., 2010; Shen
2 h at reflux temperature with agitation. As a result of esterification, et al., 2013). The mesoporous carbon–based solid acid catalysts (Csac)
90.72%, 93.18%, 91.54%, 92.95%, 91.71%, and 94.32% of lauric, have a wide range of textural properties and a relatively high density of
–SO3H groups, enabling successful conversion of FFA to esters (Malins

4285
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Table 6 due to leaching (Lou et al. 2012). According to many reports, carbon­
Catalytic performance of CCB250–SO3H catalyst on esterification of free fatty –based acid catalysts lose 30–40% of their activity after the second cycle
acids. (Konwar et al., 2014). Considering that all sulfur lost from the catalyst
Free fatty CCB250-SO3H Ethanol: FFA (molar Conversion (2.02 mmol/g) is transferred to ethyl oleate, the sulfur concentration in
acids (mmol) ratio) (%) the product is less than 0.01%, which is comparable to the Europe
Oleic acid 150 12:1 94.32 Union’s biodiesel standard (EN14214, S < 10 mg/kg). In this regard, the
Lauric acid 155 14:1 90.72 mesoporous CCB250–SO3H catalyst material presented here has great
Myristic acid 160 15:1 93.18 potential as an acidic catalyst that is stable and highly active. According
Palmitic acid 165 16:1 91.54
to Akinfalabi et al. (2017), they could only reuse their catalyst six times
Stearic acid 180 18:1 92.95
Linoleic acid 180 18:1 91.71 for biodiesel production when they used sulfonated palm seed cake
catalysts. Among the reported heterogeneous catalysts for biomass
Reaction conditions: reflux temperature at 2 h with 100 rpm agitation speed.
valorization processes, carbons (Xiong et al., 2017), zeolites (Lima et al.,
2019), heteropolyacids (Wang, Yang, 2015), and metal–organic frame­
et al., 2015; Konwar et al., 2015). works (Aljammal et al., 2019) can be found. The group of activated
carbons generated from low–cost precursors appears to be the most
versatile materials for the preparation of biomass valorization catalysts.
3.5. CCB250–SO3H catalyst reusability
Ngaosuwan et al. (2016) reported the successful synthesis of a green
sulfonated carbon–based catalyst through sulfonation of incompletely
The CCB250–SO3H catalyst developed in this study was recycled by
carbonized coffee residue (SCAC catalyst). The carbonization
centrifugation and reused for the esterification of ethanol with oleic acid
to demonstrate reusability (Fig.10). The catalyst was washed with
n–hexane and butanol after each use, and dried at 60 ◦ C overnight in a
vacuum oven. The catalyst was reused eight times at 80 ◦ C with an
ethanol–to–oleic acid molar ratio of 12 and a CCB250–SO3H catalyst
dosage of 150 mg for 2 h. In the first four cycles, oleic acid conversion
decreased from 94.32% to 91.35%. Further reduction in oleic acid
conversion was observed between cycles four and eight, 86.26%. After
eight cycles, the catalytic activity lost 8.54% after an increased number
of cycles, resulting in a decrease in ethyl oleate yield. Catalytic activity
had been reduced by water formation, although the formation of ethyl
oleate on the surface might also be a contributing factor (Fraile et al.,
2014). Zeng et al. (2014) studied a new type of sulfonated carbons
derived from the incomplete carbonization of simple natural products
such as sugar, starch, or cellulose, which has been reported to exhibit
better catalytic performance for the esterification of fatty acids and
Fig.10. Catalyst reusability: ethyl oleate yield (%) at 12:1 ethanol: oleic acid
higher stability compared to sulfonated mesoporous silica catalyst. (molar ratio), 150 mg catalyst, 80 ◦ C, 2 h with agitation.
The FT–IR spectral comparison of fresh and reused CCB250–SO3H
catalyst after the eight cycle showed that there was no significant change
in the structure of the CCB250–SO3H catalyst (Fig. 11). The presence of
–SO3H groups in the reused catalyst clearly indicates the presence of
characteristic bands (supporting information) for –SO3H and S– –O
(–SO3H stretching at 1341.50, 783.79 and 540.12 cm–1 and S– –O
stretching in –SO3H at 1010.2 cm–1). The FT–IR band of reused
CCB250–SO3H catalyst at 1250 cm–1 indicates the presence of the
O––S– –O group in the structure. It has been reported that some sulfur in
the CCB250–SO3H catalyst was lost when it was used as a catalyst for the
production of ethyl oleate (Zhou et al., 2016). After eight successive
cycles, the spent catalyst preserved 78.55% of –SO3H groups, consistent
with its sulfur content (2.02 mmol/g). Savaliya et al. (2015) reported
that approximately 87% of the original performance of SAC–SCB syn­
thesized via SCS was retained after four cycles. As determined of acid
density CCB250–SO3H of recycled catalysts decreased after eight cycles,
demonstrating the reusability of the –SO3H acid sites (Table 8). On
catalyst surfaces, unstable –SO3H functional groups leached directly,
resulting in a decrease in their density (Shen et al., 2018). After eight
cycles of reuse, the catalytic activity decreased by about 12%; the
Fig. 11. FT–IR spectroscopic analysis of (a) reused CCB250–SO3H and (b) fresh
catalyst was washed with n–hexane; elemental analysis of the reused
CCB250–SO3H catalyst.
catalyst found that sulfur content had decreased from 3.39% to 3.28%

Table 7
Comparison of the performance of various sulfonic acid treated biochar catalysts with that of sulphonated CCB250–SO3H catalyst for ethyl oleate production.
Catalysts Temperature (0C) Ethanol: oleic acid (molar ratio) Catalyst loading (wt) Reaction time (h) Ethyl oleate (%) References

CCB250–SO3H 80 12:1 150 mg 2h 94.32 This work


Rice Husk/SO3H 110 MeOH: Oil (20:1) 5g 15 h 88 Li et al., 2014
Bamboo–SO3H 90 Ethanol: oleic acid (7:1) 0.6 g 2h 91 Zhou et al., 2016
OMC–150–SO3H 80 Ethanol: oleic acid (0.5:0.005) 50 mg 10 h 73.59 Liu et al., 2008a,b

4286
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Table 8 Appendix A. Supporting information


Effect of reusability on acid value of ethyl oleate and CCB250–SO3H catalyst.
Runs Ethyl oleate Ethyl oleate acid value CCB250–SO3H acid density Supplementary data associated with this article can be found in the
yield (%) (mmol/g) (mmol/g) online version at doi:10.1016/j.egyr.2024.03.015.
1 94.32 2.89 9.426
2 94.02 2.59 9.402 References
3 93.51 2.19 9.008
4 92.85 1.94 8.925 Abdullaha, S.H.Y.S., Hanapia, N.H.M., Azida, A., Umara, R., Juahira, H., Khatoon, H.,
5 91.06 1.38 8.542 Endut, A., 2017. A review of biomass-derived heterogeneous catalyst for a
6 89.14 1.09 8.059 sustainable biodiesel production. Renew. Sustain. Energy Rev. 70, 1040–1051.
7 87.74 0.95 7.935 https://doi.org/10.1016/j.rser.2016.12.008.
8 86.36 0.81 7.404 Akinfalabi, S.I., Rashid, U., Yunus, R., Taufiq-Yap, Y.H., 2017. Synthesis of biodiesel from
palm fatty acid distillate using sulfonated palm seed cake catalyst. Renew. Energy
111, 611–619. https://doi.org/10.1016/j.renene.2017.04.056.
Aljammal, N., Jabbour, C., Thybaut, J.W., Demeestere, K., Verpoort, F., Heynderickx, P.
temperature (600 ◦ C for 4 h) and sulfonation (200 ◦ C for 18 h) were M., 2019. Metal-organic frameworks as catalysts for sugar conversion into platform
investigated, and the catalytic activity was tested via esterification of chemicals: state-of-the-art and prospects. Coord. Chem. Rev. 401, 213064 https://
caprylic acid. The catalyst was reused only for five cycles. doi.org/10.1016/j.ccr.2019.213064.
Arancon, R.A., Barros Jr, H.R., Balu, A.M., Vargas, C., Luque, R., 2011. Valorisation of
corncob residues to functionalised porous carbonaceous materials for the
4. Conclusion simultaneous esterification/transesterification of waste oils. Green. Chem. 13,
3162–3167. https://doi.org/10.1039/C1GC15908A.
Bastos, R.R.C., Correa, A.P.L., Luz, P.T.S., Rocha Filho, G.N., Zamian, J.R., Conceic Ao, L.
The CCB can be converted into a CCB250–SO3H catalyst by calcina­ R.V., 2020. Optimization of biodiesel production using sulfonated carbon-based
tion and sulfonation processes. Synthesized CCB250–SO3H catalyst catalyst from an amazon agro-industrial waste. Energy Convers. Manag. 205,
exhibited high catalytic activity due to high acid densities. We optimized 112457 https://doi.org/10.1016/j.enconman.2019.112457.
Betiku, E., Etim, A.O., Pereao, O., Ojumu, T.V., T. V, 2017. Two-step conversion of Neem
the catalyst synthesis process, activated carbon CCB250 amount, sulfo­
(Azadirachta indica) seed oil into fatty methyl esters using a heterogeneous biomass-
nation volume, sulfonation temperature, and sulfonation time had a based catalyst: an example of Cocoa pod husk. Energy Fuels. https://doi.org/
major impact on CCB250–SO3H catalyst density under optimal condi­ 10.1021/acs.energyfuels.7b00604.
Boro, J., Deka, D., Thakur, A.J., 2012. A review on solid oxide derived from waste shells
tions. It was shown that the CCB250–SO3H catalyst demonstrated high
as catalyst for biodiesel production. Renew. Sustain. Energy Rev. 16, 904–910.
catalytic activity during the esterification reaction due to its good https://doi.org/10.1016/j.rser.2011.09.011.
thermal stability, sufficient SBET of 306 m2 g–1, and pore volume of Bora, A.P., Dhawane, S.H., Anupam, K., Halder, G., 2018. Biodiesel synthesis from Mesua
1.02 cm3 /g. The CCB250–SO3H catalyst loading of 150 mg resulted in a ferrea oil using waste shell derived carbon catalyst. Renew. Energy 121, 195–204.
https://doi.org/10.1016/j.renene.2018.01.036.
94.32% oleic acid conversion with an ethanol–to–oleic acid molar ratio Cagnon, B., Py, X., Guillot, A., Stoeckli, F., Chambat, G., 2009. Contributions of
of 12:1, a 2 h reaction time, an 80 ◦ C reaction temperature, and a 2 h hemicellulose, cellulose and lignin to the mass and the porous properties of chars
reaction time. As a result of the stable attachment of the –SO3H func­ and steam activated carbons from various lignocellulosic precursors. Bioresour.
Technol. 100 (1), 292–298. https://doi.org/10.1016/j.biortech.2008.06.009.
tional group, CCB250–SO3H catalyst was successfully reused 8 times for Cao, Y., Liu, Y., Li, Z., Zong, P., Hou, J., Zhang, Q., Gou, X., 2022. Synergistic effect,
esterification cycles. For the first four reactions, the sulfur content kinetics, and pollutant emission characteristics of co-combustion of polymer-
leached was insignificant. For the esterification of higher fatty acids, the containing oily sludge and cornstalk using TGA and fixed-bed reactor. Renew.
Energy 185, 748–758. https://doi.org/10.1016/j.renene.2021.12.039.
material exhibits remarkable catalytic performances and shows minimal Chua, S.Y., Periasamy, L.A.P., Goh, C.M.H., Tan, Y.H., Mubarak, N.M., Kansedo, J.,
leaching of –SO3H groups during the reaction. As a result, producing a Khalid, M., Walvekar, R., Abdullah, E.C., 2020. Biodiesel synthesis using natural
catalyst from corn cob has several advantages, including being low cost, solid catalyst derived from biomass waste — a review. J. Ind. Eng. Chem. 81, 41–60.
https://doi.org/10.1016/j.jiec.2019.09.022.
easily synthesized, and being derived from widely available biomass
Dhawane, S.H., Kumar, T., Halder, G., 2016a. Parametric effects and optimization on
waste, making it a sustainable and efficient catalyst. This study provides synthesis of iron (II) doped carbonaceous catalyst for the production of biodiesel.
an alternative approach to the use of catalyst with high acid values in Energy Convers. Manag. 122, 310–320. https://doi.org/10.1016/j.
enconman.2016.06.005.
biodiesel production. As a solid acid catalyst, this mesoporous material
Dhawane, S.H., Kumar, T., Halder, G., 2016b. Biodiesel synthesis from Hevea brasiliensis
has the potential to be both stable and highly active and may find oil employing carbon supported heterogeneous catalyst: optimization by Taguchi
wide–ranging applications in other acid–catalyzed processes. method. Renew. Energy 89, 506–514. https://doi.org/10.1016/j.
renene.2015.12.027.
Etim, A.O., Betiku, E., Ajala, S.O., Olaniyi, P.J., Ojumu, T.V., 2018. Potential of ripe
CRediT authorship contribution statement plantain fruit peels as an ecofriendly catalyst for biodiesel synthesis: optimization by
artificial neural network integrated with genetic algorithm. Sustainability 10 (3),
Jinxi Wang: Supervision, Software, Resources, Project administra­ 707. https://doi.org/10.3390/su10030707.
Ezebor, F., Khairuddea, M., Abdullah, A.Z., Boey, P.L., 2014. Oil palm trunk and
tion. Cunliang Zhao: Writing – review & editing, Supervision, Project sugarcane baggase derived solid acid catalysts for rapid esterification of fatty acids
administration, Methodology. Bangjun Liu: Software, Resources, Proj­ and moisture-assisted transesterification of oils under pseudo-infinite methanol.
ect administration, Methodology. Kai Bian: Software, Investigation, Bioresour. Technol. 157, 254–262. https://doi.org/10.1016/j.biortech.2014.01.110.
Farabia, M.S.A., Ibrahimd, M.L., Rashide, U., Taufiq-Yap, Y.H., 2019. Esterification of
Funding acquisition, Formal analysis. Balaji Panchal: Writing – review palm fatty acid distillate using sulfonated carbon-based catalyst derived from palm
& editing, Writing – original draft, Data curation. Yuzhuang Sun: kernel shell and bamboo. Energy Convers. Manag. 181 562–570. https://doi.org/
Writing – review & editing, Visualization, Validation, Supervision, 10.1016/j.enconman.2018.12.033.
Fraile, J.M., Garcia-Bordeje, E., Pires, E., Roldan, L., 2014. New insights into the strength
Methodology. Qiaojing Zhao: Software, Resources, Project adminis­
and accessibility of acid sites of sulfonated hydrothermal carbon. Carbon 77,
tration, Formal analysis. 1157–1167. https://doi.org/10.1016/j.carbon.2014.06.059.
Fu, W., Zhang, K., Chen, M.-S., Zhang, M., Shen, Z., 2021. One-pot synthesis of N-doped
hierarchical porous carbon for high-performance aqueous capacitors in a wide pH
Declaration of Competing Interest
range. J. Power Sou. 491, 229587 https://doi.org/10.1016/j.
jpowsour.2021.229587.
There are no conflicts of interest to declare. Gohain, M., Devi, A., Deka, D., 2017. Musa balbisiana Colla peel as highly effective
renewable heterogeneous base catalyst for biodiesel production. Ind. Crops Prod.
109, 8–18. https://doi.org/10.1016/j.indcrop.2017.08.006.
Data Availability Guo, F., Xiu, Z., Liang, Z., 2012. Synthesis of biodiesel from acidified soybean soapstock
using a lignin-derived carbonaceous catalyst. Appl. Energy 98, 47–52. https://doi.
Data will be made available on request. org/10.1016/j.apenergy.2012.02.071.
Hamza, M., Ayoub, M., Shamsuddin, R.B., Mukhatar, A., Saqib, S., Zahid, I., Ameen, M.,
Ullah, S., Al-Sehemi, A.G., Ibrahim, M., 2020. A review on the waste biomass derived

4287
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

catalysts for biodiesel production. Environ. Technol. Innov. 21, 1012200. https:// Mendonça, I.M., Paes, O.A., Maia, P.J., Souza, M.P., Almeida, R.A., Silva, C.C.,
doi.org/10.1016/j.eti.2020.101200. Duvoisin, S., Freitas, F.A., 2019. New heterogeneous catalyst for biodiesel
Han, S., Yang, J., Huang, H., 2022. Novel self-solidifying double-site acidic ionic liquid as production from waste tucuma peels (Astrocaryum aculeatum Meyer): Parameters
efficient and reusable catalyst for green biodiesel synthesis. Fuel 315, 122815. optimization study. Renew. Energy 130, 103–110. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.fuel.2021.122815. renene.2018.06.059.
Hara, M., Yoshida, T., Takagaki, A., Takata, T., Kondo, J.N., Hayashi, S., Domen, K., Miranda, M.T., Sepúlveda, F.J., Arranz, J.I., Montero, I., Rojas, C.V., 2018. Analysis of
2010. Biomass conversion by a solid acid catalyst. Energy Environ. Sci. 3, 601–607. pelletizing from corn cob waste. J. Environ. Manag. 228, 303–311. https://doi.org/
https://doi.org/10.1039/B922917E. 10.1016/j.jenvman.2018.08.105.
Inagaki, M., 2009. Pores in carbon materials-importance of their control. New. Carbon Mumtaz, M., Baqar, Z., Hussain, N., Afifa Bilal, M., Azam, H.M.H., Baqir, Q.-U.-A.,
Mater. 24, 193–222. https://doi.org/10.1016/S1872-5805(08)60048-7. Iqbal, H.M.N., 2022. Application of nanomaterials for enhanced production of
Kang, S., Ye, J., Chang, J., 2013. Recent advances in carbon-based sulfonated catalyst: biodiesel, biooil, biogas, bioethanol, and biohydrogen via lignocellulosic biomass
preparation and application. Int. Rev. Chem. Eng. 5 (3), 133–144. transformation. Fuel 315, 122840. https://doi.org/10.1016/j.fuel.2021.122840.
Karmakar, A., Karmakar, S., Mukherjee, S., 2010. Properties of various plants and Naeema, M.M., Al-Sakkarib, E.G., Boffito, D.C., Gadalla, M.A., Ashour, F.H., 2021. One-
animals feedstocks for biodiesel production. Bioresour. Technol. 10, 7201–7210. pot conversion of highly acidic waste cooking oil into biodiesel over a novel bio-
https://doi.org/10.1016/j.biortech.2010.04.079. based bi-functional catalyst. Fuel 283, 118914. https://doi.org/10.1016/j.
Kastner, J.R., Miller, J., Geller, D.P., Locklin, J., Keith, L.H., Johnson, T., 2012. Catalytic fuel.2020.118914.
esterification of fatty acids using solid acid catalysts generated from biochar and Nair, A.S., Sivakumar, N., 2022. Enhanced production of biodiesel by Rhodosporidium
activated carbon. Catal. Today 190, 122–132. https://doi.org/10.1016/j. toruloides using waste office paper hydrolysate as feedstock: optimization and
cattod.2012.02.006. characterization. Fuel 327, 125174. https://doi.org/10.1016/j.fuel.2022.125174.
Kitano, M., Arai, K., Kodama, A., Kousaka, T., Nakajima, K., Hayashi, S., Hara, M., 2009. Nath, B., Das, B., Kalita, P., Basumatary, S., 2019. Waste to value addition: utilization of
Preparation of a sulfonated porous carbon catalyst with high specific surface area. waste Brassica nigra plant derived novel green heterogeneous base catalyst for
Catal. Lett. 131, 242–249. https://doi.org/10.1007/s10562-009-0062-4. effective synthesis of biodiesel. J. Clean. Prod. 239, 118112 https://doi.org/
Konwar, L.J., Das, R., Thakur, A.J., Salminen, E., Maki-Arvela, P., Kumar, N., Mikkola, J.- 10.1016/j.jclepro.2019.118112.
P., Deka, D., 2014. Biodiesel production from acid oils using sulfonated carbon Nda-Umar, U.I., Ramli, I., Muhamad, E.N., Taufiq-Yap, Y.H., Azri, N., 2022. Synthesis
catalyst derived from oil-cake waste. J. Mol. Catal. A: Chem. 388-389, 167–176. and characterization of sulfonated carbon catalysts derived from biomass waste and
https://doi.org/10.1016/j.molcata.2013.09.031. its evaluation in glycerol acetylation. Biomass. Conv. Bioref. 12, 2045–2060. https://
Konwar, L.J., Maki-Arvela, P., Salminenb, E., Kumar, N., Thakur, A.J., Mikkola, J.-P., doi.org/10.1007/s13399-020-00784-0.
Deka, D., 2015. Towards carbon efficient biorefining: Multifunctional mesoporous Ngaosuwan, K., Goodwin Jr, J.G., Prasertdham, P., 2016. A green sulfonated carbon-
solid acids obtained from biodiesel production wastes for biomass conversion. Appl. based catalyst derived from coffee residue for esterification. Renew. Energy 86,
Catal. B: Env 176, 20–35. https://doi.org/10.1016/j.apcatb.2015.03.005. 262–269. https://doi.org/10.1016/j.renene.2015.08.010.
Li, Q., Chen, S., Zhuang, L., Xu, X., Li, H., 2012. Preparation of a sulfonated activated Oginni, O., Singh, K., 2019. Pyrolysis characteristics of Arundo donax harvested from a
carbon fibre catalyst with γ-irradiation induced grafting method. J. Mater., Res. 27, reclaimed mine land. Ind. Crops Produ. 133, 44–53. https://doi.org/10.1016/j.
3083–3089. https://doi.org/10.1557/jmr.2012.360. indcrop.2019.03.014.
Li, M., Zheng, Y., Chen, Y., Zhu, X., 2014. Biodiesel production from waste cooking oil Park, Y.-M., Lee, D.-W., Kim, D.-K., Lee, J.-S., Lee, K.-Y., 2008. The heterogeneous
using a heterogeneous catalyst from pyrolyzed rice husk. Bioresour. Technol. 154, catalyst system for the continuous conversion of free fatty acids in used vegetable
345–348. https://doi.org/10.1016/j.biortech.2013.12.070. oils for the production of biodiesel. Catal. Today 131, 238–243. https://doi.org/
Lima, C.G., Jorge, E.Y.C., Batinga, L.G.S., Lima, T.M., Paixao, M.W., 2019. ZSM-5 zeolite 10.1016/j.cattod.2007.10.052.
as a promising catalyst for the preparation and upgrading of lignocellulosic biomass- Poveda-Giraldo, J.A., Solarte-Toro, J.C., Cardona Alzate, C.A., 2021. The potential use of
derived chemicals. Curr. Opin. Green. Sustain. Chem. 15, 13–19. https://doi.org/ lignin as a platform product in biorefineries: a review. Renew. Sustain. Energy Rev.
10.1016/j.cogsc.2018.08.001. 138, 110688 https://doi.org/10.1016/j.rser.2020.110688.
Lin, Q., Zhang, C., Wang, X., Cheng, B., Mai, N., Ren, J., J, 2019. Impact of activation on Prabhavathi Devi, B.L.A., Vijai Kumar Reddy, T., Vijaya Lakshmi, K., Prasad, R.B.N., R. B.
properties of carbon-based solid acid catalysts for the hydrothermal conversion of N, 2014. A green recyclable SO3H-carbon catalyst derived from glycerol for the
xylose and hemicelluloses. Catal. Today 319, 31–40. https://doi.org/10.1016/j. production of biodiesel from FFA-containing karanja (Pongamia glabra) oil in a
cattod.2018.03.070. single step, 370–337 Bioresour. Technol. 153. https://doi.org/10.1016/j.
Liu, R., Wang, X., Zhao, X., Feng, P., 2008a. Sulfonated ordered mesoporous carbon for biortech.2013.12.002.
catalytic preparation of biodiesel. Carbon 46, 1664–1669. https://doi.org/10.1016/ Ramadoss, G., Muthukumar, K., 2016. Mechanistic study on ultrasound assisted
j.carbon.2008.07.016. pretreatment of sugarcane bagasse using metal salt with hydrogen peroxide for
Liu, X., He, H., Wang, Y., Zhu, S., Ziao, X., 2008b. Transesterification of soybean oil to bioethanol production. Ultrason. Sonochem. 28, 207–217. https://doi.org/10.1016/
biodiesel using CaO as a solid base catalyst. Fuel 87, 216–221. https://doi.org/ j.ultsonch.2015.07.006.
10.1016/j.fuel.2007.04.013. Rao, B.V.S.K., Mouli, K.C., Rambabu, N., Dalai, A.K., Prasad, R.B.N., 2011. Carbon-based
Liu, T.T., Li, Z.L., Li, W., Shi, C.J., Wang, Y., 2013. Preparation and characterization of solid acid catalyst from de-oiled canola meal for biodiesel production. Catal.
biomass carbon-based solid acid catalyst for the esterification of oleic acid with Commun. 14, 20–26. https://doi.org/10.1016/j.catcom.2011.07.011.
methanol. Bioresour. Technol. 133, 618–621. https://doi.org/10.1016/j. Santos, E.M., Teixeira, A.P.D.C., da Silva, F.G., Cibaka, T.E., Araujo, M.H., Oliveira, W.X.
biortech.2013.01.163. C., Medeiros, F., Brasil, A.N., de Oliveira, L.S., Lago, R.M., 2015. New heterogeneous
Liu, Z., 2020. Comparison of hydrochar- and pyrochar-based solid acid catalysts from catalyst for the esterification of fatty acid produced by surface aromatization/
cornstalk: Physiochemical properties, catalytic activity and deactivation behavior. sulfonation of oilseed cake. Fuel 150 (2015) 408–414. Fuel. 150, 408–414. htt
Bioresour. Technol. 297, 122477 https://doi.org/10.1016/j.biortech.2019.122477. ps://doi.org/10.1016/j.fuel.2015.02.027.
Lou, W.Y., Guo, Q., Chen, W.J., Zong, M.H., Wu, H., Smith, T.J., 2012. A highly active Savaliya, M.L., Dholakiya, B.Z., B. Z, 2015. A simpler and highly efficient protocol for the
bagasse-derived solid acid catalyst with properties suitable for production of preparation of biodiesel from soap stock oil using a BBSA catalyst, 74416–74424
biodiesel. Chem. Sus. Chem. 5, 1533–1541. https://doi.org/10.1002/ RSC Adv. 5. https://doi.org/10.1039/C5RA13422F.
cssc.201100811. Sharma, M., Khan, A.A., Puri, S.K., Tuli, D.K., 2012. Wood ash as a potential
Lu, Y., Lu, Y.C., Hu, H.Q., Xie, F.J., Wei, X.Y., Fan, X., 2017. Structural characterization heterogeneous catalyst for biodiesel synthesis. Biomass. Bioenergy. 41, 94–106.
of lignin and its degradation products with spectroscopic methods. J. Spectrosc. https://doi.org/10.1016/j.biombioe.2012.02.017.
2017, 15. https://doi.org/10.1155/2017/8951658. Shen, S., Wang, C., Cai, B., Li, H., Han, Y., Wang, T., Qin, H., 2013. Heterogeneous
Lu, B., Zhou, J., Song, Y., Wang, H., Xiao, W., Wang, D., 2016. Molten-salt treatment of hydrolysis of cellulose into glucose over phenolic residue-derived solid acid. Fuel
waste biomass for preparation of carbon with enhanced capacitive properties and 113, 644–649. https://doi.org/10.1016/j.fuel.2013.06.021.
electrocatalytic activity towards oxygen reduction. Faraday Discuss. 147, 147–159. Shen, F., Guo, T., Bai, C., Qiu, M., Qi, X., 2018. Hydrolysis of cellulose with one-pot
https://doi.org/10.1039/C5FD00215J. synthesized sulfonated carbonaceous solid acid. Fuel. Process. Technol. 169,
Majamo, S.L., Amibo, T.A., Bedru, T.K., 2023. Synthesis and application of biomass- 244–247. https://doi.org/10.1016/j.fuproc.2017.10.015.
derived magnetic biochar catalyst for simultaneous esterifcation and trans- Shuit, S.H., Yee, K.F., Lee, K.T., Subhash, B., Tan, S.H., 2013. Evolution towards the
esterifcation of waste cooking oil into biodiesel: modeling and optimization. Mate. utilisation of functionalised carbon nanotubes as a new generation catalyst support
Rene. Sust. Energ. 12, 147–158. https://doi.org/10.1007/s40243-023-00236-5. in biodiesel production: an overview. Rsc. Adv. 3, 9070–9094. https://doi.org/
Malins, K., Brinks, J., Kampars, V., Malina, I., I, 2016. Esterification of rapeseed oil fatty 10.1039/C3RA22945.
acids using a carbon-based heterogeneous acid catalyst derived from cellulose. Appl. Stefanidis, S.D., Karakoulia, S.A., Kalogiannis, K.G., Iliopoulou, E.F., Delimitis, A.,
Catal. A: Gene. 519, 99–106. https://doi.org/10.1016/j.apcata.2016.03.020. Yiannoulakis, H., Zampetakis, T., Lappas, A.A., Triantafyllidis, K.S., 2016. Natural
Malins, K., Kampars, V., Brinks, J., Neibolte, I., Murnieks, R., 2015. Synthesis of activated magnesium oxide (MgO) catalysts: a cost-effective sustainable alternative to acid
carbon based heterogenous acid catalyst for biodiesel preparation. Appl. Catal. B: zeolites for the in situ upgrading of biomass fast pyrolysis oil. Appl. Catal. B:
Env. 176-177, 553–558. https://doi.org/10.1016/j.apcatb.2015.04.043. Environ. 196, 155–173. https://doi.org/10.1016/j.apcatb.2016.05.031.
Mardhiah, H.H., Ong, H.C., Masjuki, H.H., Lim, S., Pang, Y.L., 2017. Investigation of Tang, X., Niu, S., 2019. Preparation of carbon-based solid acid with large surface area to
carbon-based solid acid catalyst from Jatropha curcas biomass in biodiesel catalyze esterification for biodiesel production. J. Ind. Eng. Chem. 69, 187–195.
production. Energy Convers. Manag. 144, 10–17. https://doi.org/10.1016/j. https://doi.org/10.1016/j.jiec.2018.09.016.
enconman.2017.04.038. Tao, M.-L., Guan, H.-Y., Wang, X.-H., Liu, Y.-C., Louh, R.-F., 2015. Fabrication of
Mares, E.K.L., Gonçalves, M.A., da Luz, P.T.S., da Rocha Filho, G.N., Zamian, J.R., da sulfonated carbon catalyst from biomass waste and its use for glycerol esterification.
Conceicao, L.R.V., 2021. Acai seed ash as a novel basic heterogeneous catalyst for Fuel. Process. Technol. 138, 355–360. https://doi.org/10.1016/j.
biodiesel synthesis: Optimization of the biodiesel production process. Fuel 299, fuproc.2015.06.021.
120887. https://doi.org/10.1016/j.fuel.2021.120887.

4288
B. Panchal et al. Energy Reports 11 (2024) 4277–4289

Vadery, V., Narayanan, B.N., Ramakrishnan, R.M., Cherikkallinmel, S.K., Sugunan, S., Zeng, D., Liu, S., Gong, W., Wang, G., Qiu, J., Tian, Y., 2013. Acid properties of solid acid
Narayanan, D.P., Sasidharan, S., 2014. Room temperature production of Jatropha from petroleum coke by chemical activation and sulfonation. Catal. Commun. 40,
biodiesel over coconut husk ash. Energy 1–7. https://doi.org/10.1016/j. 5–8. https://doi.org/10.1016/j.catcom.2013.05.018.
energy.2014.04.045. Zeng, D., Zhang, Q., Chen, S., Liu, S., Wang, G., 2016. Synthesis porous carbon-based
Volpe, M., Messineo, A., Makela, M., Barre, M.R., Volpe, R., Corrado, C., Fiori, L., 2020. solid acid from rice husk for esterification of fatty acids. Micro Mesopor. Mater. 219,
Reactivity of cellulose during hydrothermal carbonization of lignocellulosic biomass. 54–58. https://doi.org/10.1016/j.micromeso.2015.07.028.
Fuel. Process. Technol. 206, 106456 https://doi.org/10.1016/j. Zeng, D., Liu, S., Gong, W., Chen, H., Wang, G., 2014. A nano-sized solid acid synthesized
fuproc.2020.106456. from rice hull ash for biodiesel production. Rsc. Adv. 4, 20535–20539. https://doi.
Wang, L., Dong, X., Jiang, H., Li, G., Zhang, M., 2014. Preparation of a novel carbon- org/10.1039/C4RA00266K.
based solid acid from cassava stillage residue and its use for the esterification of free Zhao, B., Collinson, M.M., 2010. Well-define hierarchical templates for multimodal
fatty acids in waste cooking oil. Bioresour. Technol. 158, 392–395. https://doi.org/ porous material fabrication. Chem. Mater. 22, 4312–4319. https://doi.org/10.1021/
10.1016/j.biortech.2014. cm101146n.
Wang, S.-S., Yang, G.-Y., 2015. Recent advances in polyoxometalate-catalyzed reactions. Zhang, Y., Tong, X., Yu, L., Meng, L., Guo, P., Xue, S., 2019. Highly efficient catalytic
Chem. Rev. 115, 4893–4962. https://doi.org/10.1021/cr500390v. valorization of biomass-derived hexoses and furfuryl alcohol in the presence of
Waqas, M., Aburiazaiza, A.S., Minadad, R., Rehan, M., Barakat, M.A., Nizami, A.S., 2018. polymer-based catalysts. Green. Energ. Environ. 4 (4), 424–431. https://doi.org/
Development of biochar as fuel and catalyst in energy recovery technologies. 10.1016/j.gee.2019.01.006.
J. Clean. Prod. 188, 477–488. https://doi.org/10.1016/j.jclepro.2018.04.017. Zhang, M., Sun, A., Meng, Y., Wang, L., Jiang, H., Li, G., 2015. Catalytic performance of
Xiong, X., Yu, I.K.M., Cao, L., Tsang, D.C.W., Zhang, S., Ok, Y., 2017. A review of biochar biomass carbon-based solid acid catalyst for esterification of free fatty acids in waste
based catalysts for chemical synthesis, biofuel production, and pollution control. cooking oil. Catal. Surv. Asia. 19, 61–67. https://doi.org/10.1007/s10563-014-
Bioresour. Technol. 246, 254–270. https://doi.org/10.1016/j.biortech.2017.06.163. 9182-y.
Yadav, G., Yadav, N., Ahmaruzzaman, Md, 2023. Advances in biomass derived low-cost Zhang, S., Tao, L., Zhang, Y., Wang, Z., Gou, G., Jiang, M., Huang, C., Zhou, Z., 2016. The
carbon catalyst for biodiesel production: preparation methods, reaction conditions, role and mechanism of K2CO3 and Fe3O4 in the preparation of magnetic peanut shell
and mechanisms. RSC Adv. 13, 23197. https://doi.org/10.1039/d3ra03561a. based activated carbon. Pow. Technol. 295, 152–160. https://doi.org/10.1016/j.
Yu, J.T., Dehkhoda, A.M., Ellis, N., 2010. Development of biochar-based catalyst for powtec.2016.03.034.
transesterification of canola oil. Energy Fuels. 25, 337–344. https://doi.org/ Zhou, Y., Niu, S., Li, J., 2016. Activity of the carbon-based heterogeneous acid catalyst
10.1021/ef100977d. derived from bamboo in esterification of oleic acid with ethanol. Energy Convers.
Yu, H., Niu, S., Lu, C., Li, J., Yang, Y., 2016. Preparation and esterification performance Manag. 114, 188–196. https://doi.org/10.1016/j.enconman.2016.02.027.
of sulfonated coal-based heterogeneous acid catalyst for methyl oleate production. Zhu, S., Xu, J., Cheng, Z., Kuang, Y., Wu, Q., Wang, B., Gao, W., Zeng, J., Li, J., Chen, K.,
Energy Convers. Manag. 126, 488–496. https://doi.org/10.1016/j. 2020. Catalytic transformation of cellulose into short rod-like cellulose nanofibers
enconman.2016.08.036. and platform chemicals over lignin-based solid acid. Appl. Catal. B: Environ. 268,
Yu, H., Niu, S., Lu, C., Li, J., Yang, Y., 2017. Sulfonated coal-based solid acid catalyst 118732 https://doi.org/10.1016/j.apcatb.2020.118732.
synthesis and esterification intensification under ultrasound irradiation. Fuel 208, Zong, M.H., Duan, Z.Q., Lou, W.Y., Smith, T.J., Wu, H., 2007. Preparation of a sugar
101–110. https://doi.org/10.1016/j.fuel.2017.06.122. catalyst and its use for highly efficient production of biodiesel. Green. Chem. 7,
Yuan, Q., Liu, S., Ma, M.-G., Ji, X.-X., Choi, S.-E., Si, C., 2021. The kinetics studies on 434–437. https://doi.org/10.1039/b615447f.
hydrolysis of hemicellulose. Front. Chem. 9, 781291 https://doi.org/10.3389/
fchem.2021.781291.

4289

You might also like