Energy Tech

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

RESEARCH ARTICLE

www.entechnol.de

Hydrocracking Rubber Seeds Oil for Biofuel Production


Using Bifunctional Sarulla-Derived Natural Zeolite–Ni and
Ni–Mo/SNZ Catalyst
Junifa Layla Sihombing,* Agus Kembaren, Herlinawati Herlinawati,
Ratna Sari Dewi, Ahmad Nasir Pulungan, Nia Veronika, Fithriyyah Karimah,
Rabiatul Adawiyah Pangaribuan, Rahayu Rahayu, and Ary Anggara Wibowo*

1. Introduction
Rubber seeds, due to their rich vegetable oil content, emerge as a promising
substrate for biofuel production. Harnessing the progressive catalytic hydro- The escalating trajectory of technological
cracking methodology, this research delves into the activation of Sarulla-derived and industrial progress underscores an
amplified demand for sustainable fuel
natural zeolite followed by Ni and Ni–Mo metal impregnation, subsequently
sources. Given the inevitable depletion of
subjecting rubber seed oil to hydrocracking at 400, 450, and 500 °C under fossil fuel reserves, the quest for renewable
atmospheric pressure at 0.1 MPa. Zeolite characteristics undergo notable energy alternatives has become paramount.
transformations postmetal impregnation. Importantly, at 500 °C, gasoline Among these alternatives, biofuels have
fraction selectivity exhibits superior efficacy, with the Ni/SNZ-A catalyst dem- distinguished themselves due to their sustain-
onstrating consistently high selectivity (>85%) across varied temperatures. ability, biodegradability, nontoxic attributes,
and mitigated exhaust emissions. Recent
Posthydrocracking reveals an amplification in n-paraffin, aromatic, olefin, and
scientific discourse has largely centered on
methyl ester profiles to 19.6%, 7.3%, 25.7%, and 17.7%, respectively, contrasting vegetable oils, due to their environmental
the diminution of cycloparaffin, carboxylic acid, and ketones to 3.6%, 0.2%, and benignity, renewability, and abundance.
2.5%, respectively. The aforementioned discoveries hold significant importance Intriguingly, rubber seed oil, which lies out-
in the realm of further investigation into the generation of gasoline fractions side the traditional vegetable oil spectrum,
through the utilization of natural zeolite catalysts and the incorporation of metals emerges as a promising biofuel precursor.[1]
Its utilization offers dual advantages: it does
with lower mass, all conducted under atmospheric pressure conditions with the
not encroach upon global food supplies and
ultimate objective of mitigating production expenses.
contains an appreciable oil composition of
40–46%, inclusive of 14–26% saturated fatty
acids and 72–86% unsaturated fatty acids.[2,3]
The conversion methodologies of vegetable oils into
biofuels can be bifurcated into transesterification and catalytic
J. L. Sihombing, A. Kembaren, H. Herlinawati, R. S. Dewi, A. N. Pulungan, hydrocracking.[4–8] The latter garners significant attention due
N. Veronika, F. Karimah, R. A. Pangaribuan, R. Rahayu to its capability to yield specific fuel fractions such as diesel, gas-
Department of Chemistry oline, and olefins.[4] Notably, the application of zeolitic catalysts
Faculty of Mathematics and Natural Sciences
in hydrocracking has seen substantial advancement. Subsadsana
Universitas Negeri Medan
Jl. Willem Iskandar Pasar V Medan Estate, Medan 20221, Indonesia and Ruangviriyachai, for instance, demonstrated the efficacy of
E-mail: junifalaylasihombing@unimed.ac.id HZSM-5 and HY zeolite catalysts in hydrocracking crude palm
A. A. Wibowo oil, yielding conversions of 37.20% and 38.57% respectively, with
School of Engineering a commendable diesel selectivity.[9] Furthermore, Li et al. har-
The Australian National University nessed the potential of Ni/SAPO-34 zeolite, achieving a striking
North Road, Acton, Australian Capital Territory 6201, Australia
E-mail: ary.wibowo@anu.edu.au 65% selectivity of alkanes during palm oil hydrocracking.[10] This
feat is attributed to the enhanced acidity and surface area realized
The ORCID identification number(s) for the author(s) of this article through Ni incorporation, fostering greater activity and gasoline
can be found under https://doi.org/10.1002/ente.202301318.
fraction selectivity during hydrocracking.[11]
© 2024 The Authors. Energy Technology published by Wiley-VCH GmbH. In the bifunctional catalyst, a pronounced catalytic activity is
This is an open access article under the terms of the Creative Commons
Attribution-NonCommercial-NoDerivs License, which permits use and underscored by the symbiotic interplay between the active sites
distribution in any medium, provided the original work is properly on the zeolite and the incorporated metal, as illustrated in
cited, the use is non-commercial and no modifications or adaptations Figure 1b. The genesis of Brønsted acid sites (BAS) arises from
are made. the induction of a local negative charge by Al present in the
DOI: 10.1002/ente.202301318 lattice, which is subsequently counterbalanced by a proton. In

Energy Technol. 2024, 2301318 2301318 (1 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Figure 1. Mechanistic insights into fatty acid conversion. a) Potential reaction pathways delineating the transformation of fatty acids into alkanes
and alkenes. b) Depiction of the active sites, encompassing BAS, LAS, and metal constituents on the catalyst. c) Hypothesized reaction mechanisms
underpinning the cracking of alkanes and alkenes.

contrast, Lewis acid sites (LAS) manifest when Al remains coor- explored hydrocracking of rubber seed oil using a Sarulla natural
dinatively unsaturated, especially when the zeolite framework zeolite (SNZ) catalyst loaded with Co and Co–Mo metals, yielding
sustains damage, such as through dealumination processes.[12] a gas product fraction of 64 wt% at 500 °C and a laudable 83 wt%
As depicted in Figure 1c, during the cracking reaction, the acid gasoline selectivity.[18] Building on this foundation, our research
sites mediate protonation of linear alkane C─C/C─H bonds, aspires to optimize the catalytic proficiency in hydrocracking rubber
transmuting them to transient carbonium ions. These ions seed oil into a biogasoline fraction. Therefore, the novelty of this
swiftly degrade into smaller alkane molecules or hydrogen via research lies in the natural zeolite-based catalyst used and its
carbenium (a process termed “protolytic cracking”). Notably, combination with 1 wt% Ni–Mo metal for the rubber seed oil hydro-
molecular hydrogen originates from hydrocarbon interactions cracking catalyst under reaction conditions under atmospheric
at the Brønsted site. Concurrently, the LAS exhibit the capability pressure. By integrating Ni and Ni–Mo metals with Sarulla natural
to extricate hydrides from alkanes. The catalytic sequence culmi- zeolite and judiciously calibrating temperature parameters under
nates with the disintegration of the emergent carbenium ion into atmospheric pressure, we aim to achieve peaks in conversion
diminutive carbenium ions and light alkene molecules, achieved and selectivity of liquid products to gasoline fractions by implement-
through a β-scission mechanism.[13] ing easier and more cost-effective production conditions.
The paradigm of bimetallic catalysis further augments
this conversion efficacy.[14] In light of this, Subsadsana et al.
elucidated that the amalgamation of metals conferred bifunc- 2. Results and Discussion
tional attributes to the catalyst, facilitating multifarious reactions
and consequently, amplifying hydrocarbon production.[15,16] 2.1. Surface Topography, Crystallographic Assessment, and N2
Complementarily, Kumar et al. evinced that CuNi/zeolite cata- Gas Adsorption Analysis
lysts outperformed their monometallic counterparts, attributing
their efficiency to elevated acidity levels that fostered active First, we utilize scanning electron microscopy (SEM) to investi-
deoxygenation sites.[17] An ancillary study by Sihombing et al. gate the catalyst’s surface morphology to assess its crystal

Energy Technol. 2024, 2301318 2301318 (2 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

homogeneity. Figure 2a delineates the surface morphology of the Table 1. Elements composition of SNZ, SNZ-a, Ni/SNZ-a, and NiMo/
pristine zeolite (SNZ), its activated counterpart (SNZ-A), SNZ-a catalysts.
Ni-impregnated SNZ (Ni/SNZ-A), and the dual Ni and Mo
impregnated SNZ (NiMo/SNZ-A). SNZ displays a terrain Element [Mass%] SNZ SNZ-A Ni/SNZ-A NiMo/SNZ-A
marked by inhomogeneous crystal grains aggregating into larger Si 15.73 23.72 19.20 15.28
formations. Upon activation with a mineral acid, the SNZ-A Al 5.64 3.45 6.57 5.71
surface transitions toward greater homogeneity, albeit still man-
O 61.58 49.08 61.34 61.19
ifested as substantial grains. Subsequent Ni and NiMo metal
impregnation, followed by calcination and oxidation procedures, Ni – – 0.88 0.17

culminates in a notably homogenized catalyst surface inter- Mo – – – 0.03


spersed with discernible gaps between zeolite particles (detailed Impurities 17.05 23.75 12.01 17.62
particle distribution available in Figure S1a–d, Supporting
Information). The average particle sizes of SNZ, SNZ-A,
Ni/SNZ-A, and NiMo/SNZ-A catalysts are 3879, 4247, 4185, peak intensities across specific domains. Table 2 provides a sum-
and 4962 μm, respectively. For further granularity, SEM–energy-
mary of the comparison of peak intensities within various 2θ
dispersive X-ray spectroscopy (EDX) was executed to discern the
regions for each catalyst. The peak characteristics of natural zeo-
metal catalyst dispersion on the zeolite surface. The percentage
lite can be observed within the range of 2 h = 9°–30° for both
of elements that make up the catalyst is summarized in Table 1.
mordenite and clinoptilolite. Specifically, the mordenite mineral
The percentage of Ni and Mo metals loaded on the catalyst sur-
type exhibits peaks at 2 h = 19.76°, 25.55°, 27.72°, and 30.23°, as
face is less than the amount loaded in the catalyst preparation
identified by JCPDS no. 11–0155. Likewise, the clinoptilolite
process with a total of 1%. Overlays of Si, Al, and Ni atomic rep-
mineral type displays peaks at 2 h = 21.84°, 23.53°, 26.34°, and
resentations are presented in Figure 2b-left and S2, Supporting
29.43°, consistent with JCPDS no. 25–1349. Notably, postacid
Information, while the amalgamation of Si, Al, Ni, and Mo
activation and calcination, SNZ-A showcased amplified intensi-
atomic visuals is depicted in Figure 2c-left and S3, Supporting
ties at 2θ = 27.72° and 30.23°, potentially signifying impurity
Information. As evidenced in Figure 2b-right and c-middle
reduction on the zeolite’s crystalline facets. The introduction
& -right, the metal catalysts are uniformly distributed over the
of Ni and Ni–Mo metals led to a general attenuation in peak
SNZ substrate.
Next, X-ray diffraction (XRD) analyses were conducted to intensities, likely reflecting their dispersion atop the zeolite sur-
ascertain the zeolite’s crystalline phases and mineral typology. face. These intensity variances directly influence the catalyst’s
The XRD, presented in Figure 2d, demonstrate fluctuations in crystallinity. Crystallinity degrees for SNZ, SNZ-A, Ni/SNZ-A,

Figure 2. Depiction of catalyst morphology and crystallinity. a) Low-magnification image captured using high-angle-annular dark-field–SEM.
b) Composite image showcasing Si, Al, and Ni atoms (right) contrasted with Ni atom distribution (left). c) Superimposed imagery illustrating Si,
Ni, and Mo atoms (right), with representations of Ni atom distribution (middle) and Mo atom distribution (left). d) Comparison of XRD of catalysts,
M = mordenite, C = clinoptilolite.

Energy Technol. 2024, 2301318 2301318 (3 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Table 2. Comparison of the intensity of several main peaks of catalysts. inducing a decline in surface density and, consequently,
crystallinity.
SNZ SNZ-A Ni/SNZ-A NiMo/SNZ-A Further, we subsequently undertook a nitrogen gas absorption
2θ Intensity 2θ Intensity 2θ Intensity 2θ Intensity analysis to delineate the surface area, pore volume, and pore
diameter of the SNZ, SNZ-A, Ni/SNZ-A, and NiMo/SNZ-A cat-
19.76 108 19.00 10 19.88 97 19.94 92
alysts (Figure 3). As presented in Table 3, there is a notable
20.10 72 20.11 91 20.44 58 20.20 49
increase in the specific surface area of the catalyst post the acid
21.84 87 22.05 91 22.00 98 22.02 98 activation procedure. This enhancement is likely attributed to the
23.53 81 23.71 80 23.67 96 23.70 93 dealumination process, which diminishes the mineral acid-
27.72 221 27.89 211 27.84 237 27.89 244 dissolved impurities, thereby yielding a more refined zeolite
30.23 39 30.43 35 30.42 50 30.41 35 surface. Tian et al. previously elucidated that HCl treatment
can extract particular quantities of Al from the zeolite framework,
significantly augmenting the specific surface area.[21] Conversely,
bifunctional catalysts exhibit a decline in this measure, a phe-
nomenon ascribed to metal dispersion within the catalyst’s pores
and NiMo/SNZ-A were quantified as 51.84%, 50.35%, 35.97%,
and on its surface.[22,23] The agglomeration and expansion of
and 30.98%, respectively. Acid-activated zeolite displayed dimin-
metal oxides during the calcination phase lead to a compromise
ished crystallinity relative to its pristine counterpart, suggesting a
crystallographic restructuring attributed to dealumination via
mineral acids. This process propels aluminum from the intrinsic
Table 3. Characterization of catalyst porosity: the specific surface area,
framework to its external environment. Al-Nayili et al. reported
pore volume, and pore diameter associated with each respective catalyst.
that Al-BEA dealumination with nitric acid has effectively
extracted the four-ring Al BAS and silanol nest.[19] Despite this,
Catalyst SBET Smicro Sexternal Vtotal Pore diameter
the intactness of the crystalline system remains, as corroborated [m2 g1] [m2 g1]a) [m2 g1]a) [cc g1] [nm]
by analogous peak patterns in the diffractogram.[20] A conspicu-
SNZ 70.15 5.44 64.71 0.14 3.42
ous crystallinity reduction was observed for bifunctional Ni/SNZ-
SNZ-A 107.3 5.47 101.8 0.19 3.39
A and NiMo/SNZ-A catalysts. The synergy of metal impregna-
tion, calcination, and oxidation possibly fosters the emergence Ni/SNZ-A 91.14 5.31 85.83 0.18 3.39
of new pore structures in the zeolite, attributed to the evacua- NiMo/SNZ-A 89.98 4.79 85.19 0.17 3.39
tion of intrinsically entrapped crystal water. Concurrently, the
impregnated metals occupy these pores and the zeolite surface, a)
t-plot method.

Figure 3. Characterization of pore properties. a–d) Nitrogen adsorption–desorption isotherm curves for SNZ (a), SNZ-A (b), Ni/SNZ-A (c), and
NiMo/SNZ-A (d) catalysts, respectively. e,f ) Pore size distribution curves for the catalysts SNZ, SNZ-A, Ni/SNZ-A, and NiMo/SNZ-A.

Energy Technol. 2024, 2301318 2301318 (4 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

in certain micropores, thereby reducing the zeolite’s specific Table 4. Compound content in feedstock.
surface area.[24]
The catalysts SNZ, SNZ-A, Ni/SNZ-A, and NiMo/SNZ-A Compound Formula RT [min] Area [%]
manifested characteristics typical of mesoporous materials, dis- Cyclooctane C8H16 2.690 1.26
playing pore diameters exceeding 2 nm. The incorporation of
Ethylbenzene C8H10 3.794 0.29
mesopores in these catalysts augments the bulk molecular diffu-
p-xylene C8H10 3.938 0.45
sion process. Beyond the enhancement of the zeolite surface, the
dealumination process further purges impurities ensconced Heptanoic acid, methyl ester C8H16O2 8.241 0.31
within the zeolite pores, amplifying the overall volume of the Cyclohexene, 1-methyl-4-(1-methylethenyl) C10H16 8.365 0.82
active zeolite pores.[25] However, post the metal deposition, both Undecane C11H24 11.130 1.17
the aggregate pore volume and mean pore diameter exhibited a 5-undecene C11H22 11.337 1.03
decrease, suggesting the distribution of Ni and Mo within the
Heptanoic acid C7H14O2 12.413 0.80
zeolite pores. The range of the catalyst pore size, extending
Benzene,2-butenyl C10H12 13.026 0.75
up to 3 nm, accommodates molecules such as triglycerides
(spanning from 1.5 nm to 12 μm), glycerol (0.6 nm), and methyl 1-dodecene C12H24 14.957 1.33
esters, including methyl oleate (2.5 nm), facilitating their hydro- Dodecane C12H26 15.336 1.10
cracking within the catalyst pores.[24,26] Nonanoic acid, methyl ester C10H20O 16.391 0.24
The adsorption–desorption isotherm delineated in Figure 3a–d Benzene, hexyl C12H18 17.757 0.51
highlights the existence of a hysteresis loop within a relative
Naphthalene, 1-methyl C11H10 19.039 0.97
pressure interval of 0.4–0.9, categorizing it as Type IV as per
2-undecanone C11H22O 19.391 0.95
the IUPAC standards.[27] This particular hysteresis loop aligns
with the H3 type, indicative of the aggregation of particles, akin Decanoic acid, methyl ester C11H22O 20.673 0.25
to plates, culminating in slit-shaped pores. The manifestation Cyclohexadecane C16H32 30.851 0.51
of these hysteresis loops on the isotherm graph underscores that 7-hexadecene C16H32 31.106 0.60
the enhanced mesoporous surface area does not stem from intra- Hexadecane C16H34 31.754 1.89
crystalline mesoporosity, but rather from the incorporation of
8-heptadecene C17H34 34.753 2.05
novel mesopores within the zeolite crystal lattice.[28–30] The con-
Heptadecane C17H36 35.484 3.53
nectivity of these mesopores to the external surface of the zeolite
crystal bolsters the diffusion rate of reactants within the micro- Hexadecanoic acid, methyl ester C17H34O2 43.035 0.40
structure, concurrently diminishing the likelihood of ancillary Hexadecanoic acid C16H32O2 44.469 0.55
cracking reactions.[31] A comparative analysis of the pore size dis- 11-octadecenoic acid, methyl ester C19H36O2 48.392 0.25
tribution across the catalysts is vividly portrayed in Figure 3e. A 2-nonadecanone C19H38O 48.558 0.61
larger dV(r) value for a specific pore size directly correlates to a
Octadecanoic acid, methyl ester C19H36O2 49.213 0.14
higher prevalence of pores of that particular size. The pores within
SNZ, SNZ-A, Ni/SNZ-A, and NiMo/SNZ-A catalysts predomi- Octadecanoic acid C18H36O2 50.454 0.25

nantly align in size, chiefly within the 15–40 Å (1.5–4 nm) bracket.

2.2. Mechanistic Framework of Catalytic Hydrocracking: of liquid products increased.[32] Mampuru et al. noted a rise
Evaluating Conversion and Selectivity in biogasoline production from hydrocracking used cooking
oil with a NiMo/alumina catalyst, escalating with temperature—
The compound content of rubber seed oil resulting from esteri- 35.7 wt% at 180 °C to 59.5 wt% at 250 °C.[33] Elevated tempera-
fication and transesterification as feedstock (see detailed GC tures induce heightened molecular kinetics, resulting in more
chromatogram in Figure S4, Supporting Information) is summa- frequent molecular collisions and optimizing the conversion
rized in Table 4. process of liquid products.
The cleavage of extended hydrocarbon chains into more con- Subsequent gas chromatography (GC) analysis segmented the
cise constituents can proceed via various mechanistic pathways, liquid product into gasoline (C6–C12), diesel (C16–C20), and
encompassing hydrogenation, hydrodeoxygenation, decarbonyla- heavy oil (C > 20) fractions. Table 5 suggests that the dominant
tion, dehydrogenation, and dehydration, as delineated in fraction resulting from each catalyst’s reactivity is indicative of its
Figure 1a. The efficacy and selectivity of the catalyst were exam- inherent selectivity. Generally, the gasoline fraction’s selectivity
ined during the hydrocracking of rubber seed oil into fuel frac- was most favorable at 500 °C across all catalysts. At this particular
tions, with temperature variations set at 400, 450, and 500 °C. temperature, the largest quantities of liquid product and gasoline
Table 5 delineates the conversion and selectivity trends for hydro- fractions are being generated. Despite the fact that the gasoline
cracking products for each respective catalyst (see detailed GC fraction exhibits a high degree of selectivity, the production of the
chromatogram in Figure S5–S7, Supporting Information). overall product is minimal at a temperature of 400 °C. This phe-
According to conversion data at these varied temperatures, opti- nomenon can be attributed to the limited interaction between the
mal reactivity was observed at 500 °C for bifunctional catalysts reactants and the catalyst at this temperature. Put differently,
and 450 °C for acid-activated catalysts. A discernible trend only a small proportion of reactants with lower mass can readily
emerged; as the reaction temperature elevated, the production move and effectively engage with the active site of the catalyst. On

Energy Technol. 2024, 2301318 2301318 (5 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Table 5. Hydrocracking yield and selectivity across catalysts and temperatures.

Catalyst Temp. [°C] Yield [wt%] Selectivity [%]


Liquid product Coke Gases Residue Gasoline (C6–C12) Diesel (C16–C20) Heavy oil (>C20)
SNZ-A 400 14.60 3.00 46.60 35.80 71.90 28.09 0.01
450 64.60 1.20 30.40 3.80 46.76 51.58 1.66
500 49.80 3.30 31.90 15.00 82.57 15.18 2.25
Ni/SNZ-A 400 2.00 0.40 26.00 71.60 96.69 3.31 0.00
450 14.20 3.40 52.20 30.20 86.52 13.23 0.25
500 44.40 1.00 40.60 14.00 86.25 13.15 0.60
NiMo/SNZ-A 400 9.78 3.40 31.26 55.56 59.38 39.98 0.64
450 19.65 4.50 33.58 40.27 58.14 40.50 1.36
500 40.77 1.90 54.43 2.90 81.87 17.19 0.94

the other hand, heavier molecules are yet to fully transition into be able to maintain the short-chain hydrocarbons formed. Even
the gaseous phase and approach the active site of the catalyst, more,the expansive external surface area, combined with the vast
resulting in a considerable amount of remaining feed residue. total pore volume and larger average pore diameter, facilitates the
It is these reactants with lighter molecules that significantly hydrocracking both intra and inter the catalyst structure.[35] The
contribute to the elevated proportion of gasoline at 400 °C. mesopores optimize accessibility of diesel molecules to the acid
Meanwhile, at a temperature of 450 °C, the greater the energy center, while gasoline components and coke precursors disperse
available for the vibration, rotation and translation processes of effortlessly outside the pore channels.[36]
the reactant molecules occur. This results in a greater possibility The NiMo/SNZ-A bimetal catalyst manifested consistent
for diffusion and adsorption of reactants on the active surface of selectivity for both gasoline and diesel fractions at 400 and
the catalyst, so that the opportunity for reactions to occur is 450 °C. However, at 500 °C, gasoline selectivity surged to over
greater and will ultimately result in a higher increase in liquid 80%. This accentuates the pivotal role of temperature in bolster-
product conversion. However, on the other hand, the selectivity ing catalytic activity, molecular movement, and the subsequent
for the gasoline fraction is relatively lower. This may be related to diffusion processes at the zeolite acid site, promoting the crack-
the large number of molecules in the gas phase that have not ing reaction. Nevertheless, intrinsic catalyst attributes like
fully moved and were adsorbed in the active site of the catalyst surface area and reduced total pore volume render the
and apparently require a longer reaction time to convert NiMo/SNZ-A catalyst’s performance somewhat inferior to the
completely. Much of it only comes out as a gas phase so that Ni/SNZ-A variant. The most recent investigation into catalytic
the gas yield increases. Apart from that, at this temperature, there cracking, aimed at generating gasoline fractions via zeolite-based
is still a lot of feed remaining as residue. catalysts and Ni–Mo metal, is concisely summarized in Table 6.
The SNZ-A catalyst demonstrated pronounced liquid product In light of these findings, in comparison to numerous preceding
conversion activity at 450 °C, though products from this temper- reports, this study demonstrates the capability to attain relatively
ature were primarily in the diesel fraction. This trend may be higher gasoline fractions. Additionally, the used catalyst is
attributed to suboptimal hydrocarbon chain cleavage at the zeo- founded on natural zeolite, containing merely 1 wt% metal
lite acid site. The extensive crystal size of SNZ-A, coupled with a content, thereby increasing cost-effectiveness. Furthermore,
less uniform surface, could hinder molecular access to the zeolite the utilization of process conditions under atmospheric pressure
acid site. Moreover, thermal contributions dominate, resulting in significantly expedites the production process, particularly if
a surge in liquid product conversion. further development is pursued.
Bifunctional catalysts, in contrast, displayed superior perfor- A comprehensive analysis of the liquid product from the
mance and selectivity. The Ni/SNZ-A catalyst boasted an Ni/SNZ-A catalyst at 500 °C was conducted using GC–MS
impressive >85% selectivity in the gasoline fraction across all (Figure S8, Supporting Information). The findings, juxtaposed
temperature parameters. An upward trend in liquid product yield with constituents present in rubber seed oil, are presented in
with temperature underscores the robust thermal stability of the Figure 4. An intriguing trend from the data reveals variances
bifunctional catalyst. The equilibrium between zeolitic acid and in compound group concentrations posthydrocracking. There’s
metal sites, paired with the prominent thermal role, augments an upsurge in n-paraffin, aromatic, olefin, and methyl ester com-
the catalyst’s efficiency in hydrocracking. The increased number pounds to 19.6%, 7.3%, 25.7%, and 17.7%, respectively, contrast-
of acid sites synergistically boosts the hydrocracking rate.[34] ing the diminution of cycloparaffin, carboxylic acid, and ketones
Within the catalyst, cracking transpires at the zeolite acid site to 3.6%, 0.2%, and 2.5%, respectively. This shift mirrors the
while hydrogenation takes place at the metal site. In addition, multifaceted reactions within hydrocracking, encompassing
Ni/SNZ-A has a more homogeneous surface morphology with carbon chain fragmentation, decarboxylation, isomerization, and
smaller and more uniform crystal shapes. These conditions will dehydrogenation processes (as depicted in Figure 1). A notable
provide a better catalyzed reaction on the catalyst surface and will gas phase yield suggests a prevailing decarboxylation reaction,

Energy Technol. 2024, 2301318 2301318 (6 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Table 6. The latest research in catalytic cracking to produce gasoline fractions.

Catalyst Feed Reactor Temp. [°C] Pressure Metal loading [wt%] Gasoline yield [%] References
Ni/h-Beta LDPE stirred stainless-steel autoclave 310 40 bar 7 68.7 Escola et al.[28]
NiMo/25Beta(37)60 A Soybean oil Fixed-bed flow 360 5 MPa MoO3 16 wt% 70 Ishihara et al.[34]
NiO 3.3
Pt/NiMo/β(37)60Alumina Soybean oil Fixed-bed flow 580 0.5 MPa MoO3 16 51 Shirasaki et al.[44]
NiO 3.3
Pt 1
Ni/ZSM-5 Linoleic acid Batch 450 1 atm 5 70 Haswin Singh et al.[45]
Ni/HY Palm oil continuous fixed bed 450 – 1 9.64 Istadi et al.[46]
NiCo/HZ Jatropha oil Batch 375 – 5 82.72 Aziz et al.[47]
Ni/SNZ Rubber seed oil Fixed- bed reactor 500 0.1 MPa 1 >85 This work
NiMo/SNZ Rubber seed oil Fixed- bed reactor 500 0.1 MPa 1 81.87 This work

27.84° from 237 to 216, and at 30.42° from 50 to 29. The


decreased intensity of this will correlate with a decrease in the
relative crystallinity of the Ni/SNZ-A catalyst after regeneration.
The catalytic performance of the regenerated catalyst was evalu-
ated in terms of the yield and selectivity of the liquid product fuel
fraction, as demonstrated in Figure 5b,c. A similar observation
was made by Sibao Liu et al. (2021), who reported a decrease in
peak intensity and zeolite cell unit parameters for the regenera-
tion catalyst, suggesting a relationship with the dealumination of
the zeolite during high-temperature regeneration, thereby result-
ing in an increased Si/Al ratio.[39]
Based on the data presented in Figure 5b, it can be observed
that the yield of the liquid product decreased after the catalyst
underwent regeneration. However, there was an increase in
the gas yield. Simultaneously, the quantities of remaining coke
and residue remained nearly unchanged. Upon further examina-
tion of the liquid product, selectivity data for the three fuel
fractions, as depicted in Figure 5c, were obtained. Both the initial
Figure 4. Constituent compound groups identified in rubber seed oil vis à
catalyst and the catalyst after regeneration exhibited a high
vis liquid products catalyzed by Ni/SNZ-A.
degree of selectivity toward the gasoline fraction. The reduction
in the gasoline fraction was minimal, at only 9%, indicating that,
characterized by the reduction of oxygen-rich compounds such in general, this catalyst retains its ability to produce a desirable
as ketones, which are released as gases. This aligns with Kim level of selectivity for the gasoline fraction. These findings are in
et al.’s findings where the Ni catalyst displayed adeptness in agreement with the work of Pan et al. (2019), who discovered that
C─C bond cleavage, liberating oxygen to yield CO2, while the oxidation regeneration process, along with reduction using
C─O bond fragmentation produced H2O.[37] Additionally, H2, can effectively restore the active sites of metal species and
cracking and dehydrogenation can instigate cyclization reactions, eliminate coke deposits that have a detrimental effect on the
culminating in aromatic compound formation. Concurrently, catalyst’s performance.[40] Therefore, the regeneration process
olefins might arise from the cracking of extended chain aromatic enables the catalyst to regain its effectiveness.
alkyls.[38]
3. Conclusion
2.3. Crystallinity and Activity of Catalyst Regeneration
The activation process coupled with the impregnation of Ni and
The regenerated used Ni/SNZ-A catalyst was then analyzed for Ni–Mo metals markedly influences the intrinsic characteristics
crystallinity, as shown in Figure 5a. Based on a comparison of the of the zeolite. Notably, dealumination induces a reduction in
diffractograms between the Ni/SNZ-A and Ni/SNZ-A regenera- crystallinity while concurrently augmenting the specific surface
tion, the diffractogram pattern of both catalysts is still the same area and overall pore volume. In contrast, the integration of Ni
but there is a decrease in intensity for most of the peaks. This can and Ni–Mo metals within the zeolite’s pores and surface culmi-
be seen in the area 2θ 19.88°, where the peak intensity decreases nates in decreased crystallinity, specific surface area, and total
from 97 to 68, at 22° from 98 to 68, at 23.67° from 96 to 59, at pore volume. An evident trend emerges wherein the production

Energy Technol. 2024, 2301318 2301318 (7 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Figure 5. Comparison of Ni/SNZ-A and Ni/SNZ-A regeneration. a) XRD diffractogram; b) yield distribution; and c) selectivity of fuel fraction.

of liquid products escalates with increasing reaction tempera- hydrogen gas, respectively. In a separate procedure for the Mo/SNZ-A cat-
ture. Bifunctional catalysts demonstrably outperform in terms alyst, 0.92 g (0.5 wt%) of (NH4)6Mo7O24.4H2O was dissolved in 500 mL of
of catalytic efficacy and selectivity. Specifically, the Ni/SNZ-A cat- deionized water and combined with 100 g of SNZ-A. This blend was sub-
jected to similar reflux, drying, oxidation, and reduction conditions. Finally,
alyst manifests an impressive selectivity, exceeding 85%, within
the NiMo/SNZ-A catalyst was produced by introducing 0.80 g (0.5 wt%) of
the gasoline fraction across all temperature gradients. The Ni(NO3).6H2O (previously dissolved in deionized water) to the Mo/SNZ-
gasoline fraction predominantly comprises alkanes, alkenes, A catalyst, following a similar sequence of processes. The overall catalyst
and their respective derivatives. In contrast, the diesel fraction preparation process is summarized simply in Figure 6.
is enriched with fatty acid constituents. Subsequent to the Catalysts’ Characterization: The crystalline attributes of the samples
hydrocracking process, there’s a discernible surge in the concen- were assessed using an X-ray diffractometer (XRD Shimadzu 6100) with
trations of n-paraffin, aromatic, olefin, and methyl ester com- Cu Kα radiation (λ = 1.54184 Å). Measurements were conducted at 40 kV
and 30 mA across a 2θ range of 7.00–70.00 degrees. Surface morphologi-
pounds, juxtaposed with a decline in cycloparaffin, carboxylic cal characteristics were evaluated using SEM (Zeiss EPOMH 10Zss) com-
acid, and ketones. These findings possess great significance plemented by energy-dispersive X-ray spectroscopy (EDX) for elemental
within the domain of additional exploration into the creation analysis. Nitrogen adsorption–desorption isotherms were generated using
of better-selectivity gasoline fractions synthesized under atmo- a Gas Sorption Analyzer (NOVA 1200e, Quantachrome) at 77 K. The
spheric pressured through the application of natural zeolite Brunauer–Emmett–Teller method was invoked to deduce the specific sur-
catalysts and the integration of Ni-Mo metals. face area. Pore volumes and diameters were extrapolated from the adsorp-
tion branch utilizing the Barret–Joyner–Halenda approach. Concurrently,
micropore surface and external surface areas were ascertained via the
t-plot methodology.
4. Experimental Section Rubber Seed Oil Synthesize: Rubber seeds were meticulously deshelled
and subsequently subjected to a drying process over a duration of 2-3 days.
Materials: The natural zeolite used in this study was sourced Once dried, the seeds were pulverized and subjected to extraction using
from North Sumatera, Indonesia, while the rubber seeds were harvested n-hexane as the solvent, maintained at 60 °C for a span of 2 h. Following
from rubber plantations within the same region. Gaseous agents extraction, the solvent was meticulously separated utilizing a rotary
such as nitrogen, oxygen, and hydrogen were procured from PT. Aneka
Gas Medan, Indonesia. Additionally, n-hexane, Ni(NO3).6H2O,
(NH4)6Mo7O24.4H2O, methanol, HCl, H3PO4, NaOH, and AgNO3 of
proanalyst (p.a) grade were obtained from Merck.
Catalysts Synthesize and Fabrications: The processing of natural zeolite
(SNZ) was based on the methodology proposed by Sihombing et al.[41–43]
The zeolite underwent a series of treatments: initially crushed, sieved,
immersed in distilled water for 24 h, and filtered. The resulting precipitate
was dried at 120 °C. Activation of the zeolite was achieved using 3 N HCl
and refluxed at 90 °C for 30 min. Postactivation, the zeolite was calcined at
500 °C under nitrogen atmosphere for 4 h to yield the activated Sarulla
natural zeolite (SNZ-A). To synthesize the Ni/SNZ-A catalyst, 1.6070 g
(1 wt%) of Ni(NO3).6H2O was dissolved in 500 mL of distilled water,
to which 100 g of SNZ-A was incorporated. The mixture underwent reflux
at 80 °C for 5 h and was then dried at 120 °C. Subsequent oxidation and
reduction processes were conducted at 500 °C for 2 h, using oxygen and Figure 6. Diagram of catalyst preparation.

Energy Technol. 2024, 2301318 2301318 (8 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Figure 7. Schematic of a fixed-bed reactor for hydrocracking process.

evaporator. The resultant rubber seed oil underwent processes of esterifi- was then utilized once more in the hydrocracking process of rubber seed
cation and transesterification, yielding fatty acid methyl ester (FAME), oil, and the resulting liquid product was subjected to characterization
which was subsequently used as a feedstock in the hydrocracking through the use of GC.
procedure.
Hydrocracking Process and Analysis: The efficacy and selectivity of the
catalyst were evaluated during the hydrocracking of rubber seed oil within
a fixed-bed reactor system (Figure 7). The hydrocracking procedure was Supporting Information
conducted using a catalyst-to-feed ratio of 1:6, under atmospheric pres-
Supporting Information is available from the Wiley Online Library or from
sure (0.1 MPa) and varied temperature conditions of 400, 450, and
the author.
500 °C. Subsequent to the process, the resultant product was channeled
through a silicone hose, directed to a condenser. The accumulated liquid
product was precisely weighed and quantified. Product yield (Y) was
calculated based on the ratio of product mass (M) to feed using Acknowledgements
Equation (1)–(4). This liquid fraction, indicative of the fuel component,
was subsequently subjected to characterization using GC and GC–mass The authors express their gratitude to the Institute for Research and
spectrometry (QP2010 Plus, Shimadzu). The selectivity of the liquid prod- Community Service (LPPM) at Universitas Negeri Medan for the financial
uct to the fuel fraction was calculated using Equation (5)–(7) (Min-Yee backing accorded under the Applied Product Research Scheme, with the
Choo et al. 2020). grant reference no. 0137/UN33.8/KPT/PPT/2023.
Open access publishing facilitated by Australian National University, as
Mliquid product part of the Wiley - Australian National University agreement via the Council
Y liquid product ð%Þ ¼  100% (1)
Mfeed of Australian University Librarians.
Mresidue
Y residue ð%Þ ¼  100% (2)
Mfeed
Mcoke
Conflict of Interest
Y coke ð%Þ ¼  100% (3)
Mfeed The authors declare no conflict of interest.
Y gas ð%Þ ¼ 100%  ½Y liquid product þ Y residue þ Y coke  (4)

Total area of hydrocarbon C6–C12


Gasoline selectivity ð%Þ ¼  100% Author Contributions
Total area of hydrocarbon
(5) J.L.S. formulated the research concept, designed the experiments, and
supervised the project. Data analysis and manuscript preparation were
Total area of hydrocarbon C16–C20 collaboratively undertaken by J.L.S., A.A.W., H.H., A.K., and A.N.P.
Diesel selectivity ð%Þ ¼  100% (6)
Total area of hydrocarbon Catalyst synthesis and execution of the hydrocracking process were led
by J.L.S., with contributions from N.V. and R.A.P. Instrumentation
Total area of hydrocarbon>C20
Heavy oil selectivity ð%Þ ¼  100% (7) arrangements and data elucidation were overseen by R.S.D., F.K., and
Total area of hydrocarbon R.R. All contributors engaged in result analysis and manuscript reviews.
Regeneration of Catalyst: The catalyst employed in the hydrocracking
procedure was subsequently subjected to a washing process utilizing
acetone in order to eliminate any impurities. In accordance with this, Data Availability Statement
the catalyst went through oxidation at 500 °C for 2 h, in the presence of
oxygen gas flow. Following that, the catalyst experienced a reduction The data that support the findings of this study are available from the
process at 500 °C for 2 h, with hydrogen gas flow. The regenerated catalyst corresponding author upon reasonable request.

Energy Technol. 2024, 2301318 2301318 (9 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH
21944296, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ente.202301318 by Nat Prov Indonesia, Wiley Online Library on [25/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.entechnol.de

Keywords [23] M. Ebrahiminejad, R. Karimzadeh, Adv. Powder Technol. 2019, 30,


1450.
bifunctional catalysts, catalytic hydrocracking, rubber seed oils, [24] M. Chareonpanich, Z.-G. Zhang, A. Tomita, Energy Fuels 1996, 10,
selectivities 927.
Received: October 20, 2023 [25] B. Wei, L. Jin, D. Wang, Y. Xiong, H. Hu, Z. Bai, Fuel 2020, 266,
Revised: January 25, 2024 117089.
Published online: [26] K. Wijaya, M. Shidiq, M. Fahrurrozi, Asian J. Chem. 2014, 26, 5033.
[27] P. P. Dik, I. G. Danilova, I. S. Golubev, M. O. Kazakov, K. A. Nadeina,
S. V. Budukva, V. Y. Pereyma, O. V. Klimov, I. P. Prosvirin,
E. Y. Gerasimov, T. O. Bok, I. V. Dobryakova, E. E. Knyazeva,
[1] W. Roschat, T. Siritanon, B. Yoosuk, T. Sudyoadsuk, V. Promarak,
I. I. Ivanova, A. S. Noskov, Fuel 2019, 237, 178.
Renewable Energy 2017, 101, 937.
[28] J. M. Escola, J. Aguado, D. P. Serrano, L. Briones, J. L. Díaz de Tuesta,
[2] E. Yousif, B. Abdullah, H. Ibraheem, J. Salimon, N. Salih, J. Al-Nahrain
R. Calvo, E. Fernandez, Energy Fuels 2012, 26, 3187.
Univ. Sci. 2013, 16, 1.
[29] Y. Li, C. Zhang, Y. Liu, S. Tang, G. Chen, R. Zhang, X. Tang, Fuel 2017,
[3] M. Yang, W. Zhu, H. Cao, Bioresour. Bioprocess. 2021, 8, 45.
189, 23.
[4] H. K. A. G. Singh, S. Yusup, C. K. Wai, Procedia Eng. 2016, 148, 426.
[30] P. Lanzafame, S. Perathoner, G. Centi, E. Heracleous, E. F. Iliopoulou,
[5] L. Li, K. Quan, J. Xu, F. Liu, S. Liu, S. Yu, C. Xie, B. Zhang, X. Ge, Fuel
K. S. Triantafyllidis, A. A. Lappas, ChemCatChem 2017, 9, 1632.
2014, 123, 189.
[31] M. O. Kazakov, K. A. Nadeina, I. G. Danilova, P. P. Dik, O. V. Klimov,
[6] J. Gimbun, S. Ali, C. C. S. C. Kanwal, L. A. Shah, N. H. M. Ghazali,
V. Y. Pereyma, E. A. Paukshtis, I. S. Golubev, I. P. Prosvirin,
C. K. Cheng, S. Nurdin, Procedia Eng. 2013, 53, 13.
E. Y. Gerasimov, I. V. Dobryakova, E. E. Knyazeva, I. I. Ivanova,
[7] H. N. T. Le, K. Imamura, N. Watanabe, M. Furuta, N. Takenaka,
A. S. Noskov, Catal. Today 2019, 329, 108.
L. Van Boi, Y. Maeda, Chem. Eng. Technol. 2018, 41, 1013.
[32] M. S. El-Sawy, S. A. Hanafi, F. Ashour, T. M. Aboul-Fotouh, Fuel 2020,
[8] S. E. Onoji, S. E. Iyuke, A. I. Igbafe, M. O. Daramola, Energy Fuels
269, 117437.
2017, 31, 6109. [33] M. B. Mampuru, D. B. Nkazi, H. E. Mukaya, Energy Sources, Part A
[9] P. Sangdara, M. Subsadsana, C. Ruangviriyachai, Orient. J. Chem. 2020, 42, 2564.
2017, 33, 2257. [34] A. Ishihara, N. Fukui, H. Nasu, T. Hashimoto, Fuel 2014, 134, 611.
[10] T. Li, J. Cheng, R. Huang, W. Yang, J. Zhou, K. Cen, Int. J. Hydrogen [35] X. Zhang, Q. Guo, B. Qin, Z. Zhang, F. Ling, W. Sun, R. Li, Catal.
Energy 2016, 41, 21883. Today 2010, 149, 212.
[11] I. T. A. Aziz, W. D. Saputri, W. Trisunaryanti, S. Sudiono, A. Syoufian, [36] X. Meng, Z. Lian, X. Wang, L. Shi, N. Liu, Fuel 2020, 270, 117426.
A. Budiman, K. Wijaya, Period. Polytech., Chem. Eng. 2021, 66, 101. [37] S. K. Kim, S. Brand, H. Lee, Y. Kim, J. Kim, Chem. Eng. J. 2013, 228,
[12] E. T. C. Vogt, B. M. Weckhuysen, Chem. Soc. Rev. 2015, 44, 7342. 114.
[13] Z. Gholami, F. Gholami, Z. Tišler, M. Tomas, M. Vakili, Energies 2021, [38] L. Marlinda, M. Al-Muttaqii, A. Roesyadi, D. H. Prajitno, J. Phys.: Conf.
14, 1089. Ser. 2020, 1442, 012048.
[14] H. He, Z. Zou, W. Hua, Y. Yue, Z. Gao, Chem. Res. Chin. Univ. 2023, [39] S. Liu, P. A. Kots, B. C. Vance, A. Danielson, D. G. Vlachos, Sci. Adv.
39, 1064. 2021, 7, eabf8283.
[15] M. Subsadsana, P. Kham-or, P. Sangdara, P. Suwannasom, [40] H. Pan, D. Xu, C. He, C. Shen, Catalysts 2018, 9, 23.
C. Ruangviriyachai, J. Fuel Chem. Technol. 2017, 45, 805. [41] J. L. Sihombing, H. Herlinawati, A. N. Pulungan, L. Simatupang,
[16] K. D. Nugrahaningtyas, N. Rahmawati, F. Rahmawati, Y. Hidayat, R. Rahayu, A. A. Wibowo, Arabian J. Chem. 2023, 16, 104707.
Open Chem. 2019, 17, 1061. [42] J. L. Sihombing, S. Gea, A. N. Pulungan, A. A.. Wibowo,
[17] R. Kumar, V. Strezov, T. Kan, H. Weldekidan, J. He, S. Jahan, Energy B. Wirjosentono, J. Phys.: Conf. Ser. 2018, 1116, 042034.
Fuels 2020, 34, 389. [43] J. L. Sihombing, A. N. Pulungan, M. Zubir, A. A. Wibowo, S. Gea,
[18] J. L. Sihombing, S. Gea, B. Wirjosentono, H. Agusnar, A. N. Pulungan, B. Wirjosentono, Y. A. Hutapea, Rasayan J. Chem. 2019, 12, 205.
H. Herlinawati, M. Yusuf, Y. A. Hutapea, Catalysts 2020, 10, 121. [44] Y. Shirasaki, H. Nasu, T. Hashimoto, A. Ishihara, Fuel Process.
[19] A. Al-Nayili, M. Albdiry, N. Salman, Arabian J. Sci. Eng. 2021, 46, 5709. Technol. 2019, 194, 106109.
[20] A. Pulungan, A. Kembaren, N. Nurfajriani, F. Syuhada, J. Sihombing, [45] H. K. Gurdeep Singh, S. Yusup, A. T. Quitain, B. Abdullah, M. Ameen,
M. Yusuf, R. Rahayu, Pol. J. Environ. Stud. 2021, 30, 5681. M. Sasaki, T. Kida, K. W. Cheah, Environ. Res. 2020, 186, 109616.
[21] H. Tian, S. Liu, Q. Liu, RSC Adv. 2022, 12, 24654. [46] I. Istadi, T. Riyanto, L. Buchori, D. D. Anggoro, A. W. S. Pakpahan,
[22] D. P. Upare, S. Park, M. S. Kim, Y.-P. Jeon, J. Kim, D. Lee, J. Lee, A. J. Pakpahan, Int. J. Renewable Energy Dev. 2021, 10, 149.
H. Chang, S. Choi, W. Choi, Y.-K. Park, C. W. Lee, J. Ind. Eng. [47] I. Aziz, D. Gustama, T. Retnaningsih, N. Saridewi, L. Adhani,
Chem. 2017, 46, 356. A. A. Dwiatmoko, M. Ridwan, Rasayan J. Chem. 2022, 15, 2026.

Energy Technol. 2024, 2301318 2301318 (10 of 10) © 2024 The Authors. Energy Technology published by Wiley-VCH GmbH

You might also like