You are on page 1of 16

Received: 7 April 2021 Revised: 25 June 2021 Accepted: 29 June 2021

DOI: 10.1002/cjce.24284

ARTICLE

Catalytic transesterification of glycerol with dimethyl


carbonate to glycerol carbonate with Co3O4 nanoparticle
incorporated MCM-41 derived from rice husk

Shivali Arora1 | Vijayalakshmi Gosu1 | Verraboina Subbaramaiah1 |


Tian C. Zhang2

1
Department of Chemical Engineering,
Malaviya National Institute of Technology
Abstract
Jaipur, Jaipur, India A rice husk-derived MCM-41 catalyst has been synthesized, and active metal
2
Department of Civil & Environmental Co3O4 nanoparticles were incorporated. The synthesized MCM-41 derived
Engineering, University of Nebraska-
from rice husk exhibited highly ordered p6mm hexagonal symmetry. The cata-
Lincoln, Omaha, Nebraska, USA
lytic activity of the prepared catalyst was investigated for glycerol carbonate
Correspondence (GLc) synthesis through transesterification of glycerol (GLy) with dimethyl
Vijayalakshmi Gosu and Verraboina
Subbaramaiah, Department of Chemical
carbonate (DMC). Experimentally, it was found that 5 wt.% cobalt incorpo-
Engineering, Malaviya National Institute rated MCM-41 activated at 400 C showed a higher yield of GLc, 60.2%, due to
of Technology Jaipur, 302017, India. its optimum basicity. Approximately, 98.7 ± 1.2% GLy conversion and 94.1
Email: vlakshmi.chem@mnit.ac.in and
vsr.chem@mnit.ac.in ± 0.7% GLc yield was achieved with the batch reactor at adjusted conditions: a
molar ratio of DMC-to-glycerol = 3, catalyst dosage = 6 wt.% (based on glyc-
Funding information
erol mass), reaction temperature = 90 C, and reaction time = 2 h. The trans-
Ministry of Education; Science and
Engineering Research Board, Grant/ esterification reaction kinetics were also studied using the ODE 15s tool in
Award Numbers: ECR/2015/000085, MATLAB. The catalyst’s activation energy was calculated as 58.2 ± 2.4 kJ/mol.
Dated 20.08.2016
KEYWORDS
glycerol carbonate, MCM-41, nanoparticles, rice husk, transesterification

1 | INTRODUCTION attention due to its high boiling point, low toxicity, and
biodegradability. It can be used as a novel solvent, surfac-
Recently, renewable fuels have received significant atten- tant, electrolyte in a lithium-ion battery, and monomer
tion to meet the present demand. Among the various for polymers. Many researchers have paid attention to
renewable fuels, biodiesel is a potential alternative GLc synthesis. The exploration of state-of-the-art technol-
source. However, every 100 kg biodiesel production gen- ogy is an important challenge for the industrial-scale syn-
erates nearly 10 kg of glycerol (GLy) as a by-product. thesis of GLc. Up to now, several methods have been
Effective utilization of this by-product GLy could adopted in the literature to make GLc using GLy as
enhance the sustainability of the biodiesel industry. feedstock with various reactants: (i) CO2, (ii) urea,
Besides, an abundant quantity of GLy may hamper mar- (iii) CO, (iv) phosgene, and (v) ethylene or dimethyl
ket prices that have dropped drastically since 2006.[1] carbonate (DMC).[2] However, the lowest conversion of
Numerous ventures have been explored for the valori- GLy was obtained with CO2, since the reaction is ther-
zation of GLy into several value-added chemicals, such modynamically limited. In the case of urea, ammonia
as, glycidol, glycerol carbonate (GLc), propanediol, and formed as a by-product, which needs to be removed
polyamides. Among them, GLc has received significant under vacuum.

Can J Chem Eng. 2021;1–16. wileyonlinelibrary.com/journal/cjce © 2021 Canadian Society for Chemical Engineering. 1
2 ARORA ET AL.

In contrast, CO and phosgene as reactant are toxic. produced, especially in India, of which around 24 MTPA
Among various synthesizing routes, the trans- of rice husk are produced by the rice milling industry.[14]
esterification of GLy with DMC is an environmentally During rice husk combustion, approximately 14–25 wt.%
benign and industrially feasible route to produce GLc.[3] of ash is generated, which contains silica as high as 90%
However, without the aid of any suitable catalyst, GLy to 97% and has been confirmed to be ideal silica due to
and DMC are immiscible, forming a biphasic mixture its significant portion of silica oxide. [15,16] Several other
with a negligible conversion of GLy.[4] agricultural sources or biomasses such as rice straw,
The transesterification reaction needs a basic site cat- sugarcane bagasse, bamboo culm, bamboo leaf, corn-
alyst to accelerate the rate of reaction.[5] Homogeneous cob, and banana leaf were also employed as silica
basic catalysts such as NaOH, KOH, K2CO3, and MgO sources.[17]
have been used; such catalysts have provided excellent Nanoparticles of metal and metal oxides have been an
results in GLc yield.[6] However, the separation and indispensable area of research for the last few decades
regeneration of catalyst from the product mixture were owing to their high-performance catalytic activity and
poor. In this view, several methods were adopted for het- good interaction with support material. Among various
erogenizing homogeneous catalysts to overcome the metal oxides, Co3O4 nanoparticles were found to be a
problems mentioned above. A heterogeneous catalyst can promising catalyst due to their small Co-O bond
be easily separated from the product mixture, reused, energy.[18] In several instances, unsupported metal oxide
and barely corrodes the reactor. Researchers have utilized nanoparticles are not stable because of the high surface-
various catalysts, such as Ca/La,[7] Mg/Al/Zr,[8] Li/ZnO,[9] area-to-volume ratio. Generally, nano-metal oxides are
and Mg/Ce[10] as basic heterogeneous catalysts for tailoring supported onto porous materials to inhibit the agglomer-
the transesterification reaction. However, these catalysts ation and to enhance the recovery. The catalytic perfor-
have certain limitations. For instance, Li/ZnO exhibits poor mance of support material enhances upon introduction
reusability due to active species leaching during the reac- of oxides of nanoparticles through synergistic effects or
tion. Similarly, hydrotalcite requires a long reaction time metal-support interactions. Hence, in the present study,
and a high DMC-to-GLy molar ratio to achieve a high Co3O4 nanoparticles are introduced upon stable support
conversion of GLy.[11] In sequence, these catalysts possess material, such as MCM-41.
low surface area due to the non-utilization of the highly In this study, rice husk was used as a silica precursor
porous catalyst support. Thus, an environmentally benign, for the synthesis of mesoporous catalysts (MCM-41). The
reusable, inexpensive, and non-toxic solid-base catalyst basic sites of the catalyst were boosted by varying active
needs to be explored for the synthesis of GLc. metal loading and calcination temperature. Batch experi-
Mesoporous catalysts incorporated with active metal ments were conducted to maximize yield and conversion
oxide species have fascinated researchers due to their by adjusting influencing conditions. The stability of the
excellent catalytic activity, high surface area, and easy to prepared catalyst was investigated through a reusability
tailor pore length and pore diameter to enhance the reac- study. The reaction’s apparent activation energy was cal-
tion. Among the various mesoporous support materials, culated with the reaction’s kinetic data using the ODE
mesoporous silica has received rapid attention in mate- 15s tool in MATLAB.
rial science and catalysis due to its uniform channels,
high surface area, adjustable pore sizes, and high thermal
stability.[6] MCM-41 is a mesoporous silica material with 2 | MATERIALS AND METHODS
a two-dimensional hexagonal array of mesopores and a
large pore size. MCM-41 can be synthesized using various 2.1 | Materials
silica sources, including aerosol, n-alkyl amines, n-
alkoxysilanes, TEOS, and water glass. However, purely Rice husk was obtained from a rice mill in Agra, India. Cobalt
siliceous MCM-41 has relatively low catalytic activity due nitrate hexahydrate (Co[NO3]2.6H2O) was acquired from
to the chemically inert silicate framework; besides, these CDH Fine Chemicals, New Delhi, India. Tetraethyl amine
silica precursors are expensive and may eventually (TEA) was purchased from Avra Chemicals, Hyderabad,
increase the process cost.[12] India. Cetyltrimethylammonium bromide (CTAB), nitric acid,
Some researchers have attempted a more economical phenolphthalein, 2, 4-nitroaniline, sodium hydroxide pellets,
approach using various waste resources: coal fly ash, rice and bromothymol blue were procured from Fisher Scientific,
husk, kaolin, red mud, and so on to make catalysts.[2,5,6] India. Methanol was supplied by Qualichem Fine Chemicals,
Utilization of waste resources is a highly motivated Mumbai, India. Ethanol was bought from Merck, India.
recycling technique which reduces waste management Glycerol, dimethyl carbonate, and acetic acid were obtained
costs.[13] Approximately 120 MTPA of rice paddy are from Loba Chemie, Mumbai, India.
ARORA ET AL. 3

2.2 | Preparation of catalyst (Massachusetts, USA) under nitrogen atmosphere with a


heating rate of 20  C/min from room temperature
In this study, the catalyst was prepared by adopting a (25  C) to 900  C. X-ray data of the bare MCM-41 and
two-step process: (i) synthesis of catalyst support material cobalt species-loaded MCM-41 was attained on PAN-
followed by (ii) impregnation of Co3O4 nanoparticles on alytical X’Pert Pro equipment. To obtain the X-ray
the synthesized support material. Initially, the catalyst diffraction (XRD) pattern, the equipment was oper-
support material, MCM-41, was prepared using rice husk ated at 40 kV, 40 mA, where Cu Kα was used as the
as a silica source.[6,19] Rice husk was cleaned with water, target source between the range of 2θ = 10–90  . In
dried, and acid-treated with 1 mol/L nitric acid to remove the same way, the small-angle powder scattering was
adhered dust and other impurities. The acid-treated measured over the 2θ region of 2–10  . The average
material was placed in a furnace at 700 C to obtain silica crystal size of a particle was calculated using the
fuse. Then, 10 g of white silica fuse was dissolved in Scherrer equation:
500 ml of sodium hydroxide solution (1 mol/L) for 18 h
to acquire a sodium silicate solution. Then, 9.8 g of sur- 0:9 λ
τ¼ ð1Þ
factant (CTAB) was dissolved in 50ml of distilled water, βcosθ
and 400 ml of sodium silicate was added dropwise into
it. Under continuous stirring, the pH of the solution was where τ is the average crystallite size (nm), λ is the X-ray
adjusted to 11 using 1 mol/L acetic acid and was con- wavelength at 1.54 nm, β is the full width at half-
stantly stirred for 6 h. The obtained solution was trans- maximum (FWHM), and θ is Bragg’s angle.[20]
ferred into a polypropylene bottle and placed into a hot A PerkinElmer Fourier transform spectrophotom-
air oven at 100 C for 72 h. The formed precipitate (gel) eter was used to record the Fourier transform infrared
was filtered and washed with ethanol (99%); the washed (FTIR) spectrum of prepared samples in the region of
gel was dried in an oven at 100 C for 24 h. The dried 4000–400 cm 1 . Before analysis, a translucent circu-
white powder was kept in a furnace at 550 C to remove lar pellet of catalyst and KBr was prepared and ana-
the surfactant material (CTAB). The calcined product lyzed. The textural properties of the catalyst were
was designated as MCM-41. studied on a Quantachrome NovaWin (Version 11.05,
In the second step, cobalt nanoparticles were pre- USA). The impurities of the catalyst were degassed at
pared using a surfactant method.[20] The synthesized 350  C for 4 h before analysis. The surface area of the
nanoparticles were then introduced on MCM-41. In a catalyst support and cobalt-loaded support material
typical procedure, cobalt nitrate hexahydrate was mixed was calculated using the BET method, whereas the
with CTAB in a molar ratio of 5, followed by heating at pore volume and diameter were calculated using the
75 C for 0.5 h. Thereafter, triethylamine (TEA) was used BJH method. The morphology of the catalyst was stud-
as a reducing agent with a molar ratio of 0.5 (Co:TEA) by ied through an FEI Nova NanoSEM 450. Energy-
dropwise addition with continuous stirring. During the dispersive X-ray spectroscopy (EDX) was done to
acquisition of TEA, dark brown particles appeared due to determine the dispersion of active metal on the sup-
the formation of cobalt nanoparticles. Then, the prepared port material. The basic strength of the prepared cata-
MCM-41 was added into the dark brown solution and lyst was calculated by adopting the Hammett
aged for 10 h under total reflux condition. The resultant indicator method with phenolphthalein (H_ = 9.3),
solution was centrifuged and filtered with a Whatman fil- 2,4-dinitroaniline (H_ = 15), and bromothymol blue
ter paper. Finally, the filtrate was washed with ethanol, (H_ = 7.8) as indicators. The method adopted for the
dried in an oven at 100 C, and then placed in a muffle quantification of basic sites was as described by Arora
furnace for calcination at 450 C for 2 h. Different samples et al.[21] Firstly, the catalyst was added to a solution
of catalysts were prepared and titled as Co/MCM-41 (x%, T), containing methanol and indicator and left for 2 h. If
where x% and T refers to the weight percentage of cobalt colour change was observed, the catalyst was consid-
loading and calcination temperature of the prepared catalyst ered as stronger than the indicator; otherwise, it was
in  C, respectively. weaker than the indicator. To quantify the number of
basic sites, the solution was titrated with benzoic acid.
The samples’ internal morphology was studied on a
2.3 | Catalyst characterization Tecnai G2 20 S-Twin (FEI, Netherlands) apparatus.
The X-ray photoelectron spectroscopy (XPS) spectra
The stability of the bare and cobalt-loaded MCM-41 cata- were recorded on a Scienta Omicron spectrometer
lyst was analyzed on a PerkinElmer STA-6000 using Al Kα source.
4 ARORA ET AL.

2.4 | Transesterification reaction of GLy GLc moles produced, GLc selectivity, time, and GLy
with dimethyl carbonate molecular mass, respectively.

The procedure adopted for the transesterification of GLy


to GLc was similar to that described by Arora et al.[21] 3 | RESULTS A ND DISCUSSION
Typically, required amount of GLy and DMC was
charged in round-bottom flask and heated until the 3.1 | Catalyst characterization
desired temperature was attained. The required catalyst
was added into the reaction mixture and the reaction was Figure 1A shows the wide-angle XRD profiles of cobalt
proceeded for 2 h. The product was quantified by gas nanoparticles, rice husk, acid treated rice husk, rice husk
chromatography using pyridine as internal solvent. In all ash, MCM-41, and Co-loaded MCM-41. The existence of
the calculations, GLy was assumed as a limiting reactant. the broad hump at 2θ = 23 in all the samples except
The GLy conversion, GLc yield, and selectivity, as well as Co3O4 and rice husk ash may be attributed to the amor-
turn over frequency (TOF), were calculated using the fol- phous nature of silica. The characteristic peaks of cobalt
lowing formulas, respectively: oxide (Co3O4) nanoparticles are located at 2θ = 19.06 ,
31.32 , 36.91 , 44.87 , 59.39 , 65.28 , and 77.39 . This
 
GLyinitial GLyunreacted illustrates its crystalline behaviour. The average crystal-
XGLy ¼  100 ð2Þ
GLyinitial line size was calculated based on the Co3O4 diffraction
peak (311), using Equation (1) and was about ~32.94 nm.
  The current work is in good agreement with Zhu
GLcproduced
YGLc ¼  100 ð3Þ et al.,[22] who calculated the crystallite size of spinel
GLyinitial
Co3O4 particles prepared from hydrothermal treatment
to be 35.4 nm.
 
YGLc Rice husk, acid-treated rice husk, and rice husk ash
SGLc ¼  100 ð4Þ samples’ XRD profiles are depicted in Figure 1A. Rice
XGLy
husk and acid-treated rice husk samples’ observed peak
  at 27.4 [23] was attributed to partially crystallized carbon
GLyinitial  XGLy  MGLy from cellulose; upon calcination, the peak disappeared.
TOF ¼  100
100  mass of catalyst used for reaction  t Rice husk ash sample (silica fuse) exhibited a broad
ð5Þ hump around 22 , characteristic of amorphous silica.[24]
Thus, it confirmed that calcination of acid-treated rice
where XGLy, GLyinitial, GLyunreacted, YGLc, GLcproduced, husk removes carbon impurities present in it and the
SGLc, t, and MGLy, represent GLy conversion, moles of carbon will not interfere during the synthesizing process
GLy in feed (initial), GLy moles in product, GLc yield, of MCM-41.

FIGURE 1 X-ray diffraction (XRD) profile: (A) wide and (B) small-angle profile
ARORA ET AL. 5

The presence of strong peaks at 36.91 , which corre- applied to decompose lignin present in the husk as it is
spond to Co3O4, justifies the incorporation of Co3O4 the most stable aromatic polymer of rice husk.[26,27]
nanoparticles into the MCM-41 framework. The strength The remaining weight of the sample after 650 C was
of the peak at 36.91 increased with an increase in the silica ash. Based upon TGA results, 700 C was selected as
loading amount of cobalt in the channels of MCM-41. calcination temperature to get the silica fuse from acid-
The small-angle X-ray scattering profile of bare MCM-41 treated rice husk, which was further used for the synthesis of
and Co/MCM-41 is displayed in Figure 1B. The planes MCM-41.
(100), (110), (200), and (210) are ascribed to the hexago- The devolatilization spectrum of MCM-41 and
nal space group p6mm.[6] The intensity of the major peak Co/MCM-41 (5%, 400 C) is presented in the inset of
at the (100) plane was noticeable even after loading a higher Figure 2A. From Figure 2A, two weight-loss zones were
percentage of cobalt oxide nanoparticles, indicating that perceived in the TGA. The first weight loss zone was
MCM-41 has a high degree of hexagonal organization in observed below ~300 C, indicating the loss of water mol-
the mesoporous region. At a higher concentration of cobalt ecules present in the pores of the MCM-41 framework.[11]
species in MCM-41, the peak strength decreased owing to Furthermore, the second weight loss zone observed in
distortion of the porous hexagonal structure of MCM-41. the range of ~300–565 C might be due to the burning of
Furthermore, the Co/MCM-41 samples’ peaks were left-over CTAB molecules. No further losses were
observed broadening compared to the parent MCM-41 due witnessed upon an increase in the temperature until
to the reduction of inter-planar d-spacing from 3.28–2.81 nm. 800 C, justifying the support material’s stability. Simi-
It may be due to the incorporation of Co3O4 nanoparticles in larly, in the spectra of Co/MCM-41 (5%, 400 C), the first
the channels of MCM-41. The present results obtained were weight loss zone was up to 300 C, similar to the parent
in good agreement with the previous reports by Sikarwar MCM-41, and occurred owing to the loss of adsorbed
et al.[6] and Huo et al.[25] water molecule in the catalyst pores. The second weight-
Thermogravimetric analysis (TGA) was performed to loss region in the range of 550–850 C might be due to the
confirm the calcination temperature of the rice husk and decomposition of metal nitrate into oxides, which is
acid-treated rice husk. TGA curves of rice husk and acid- responsible for accelerating the transesterification
treated rice husk showed a similar pattern (Figure 2A). reaction.
Thermal decomposition of hemicellulose and cellulose The FTIR spectra of washed rice husk, acid-treated
was observed with a temperature range of 250–370 C. rice husk, rice husk ash, MCM-41, and cobalt impreg-
Firstly, hemicellulose was decomposed and then cellulose nated MCM-41 in the fingerprint region of 4000–
decomposed in temperature ranging from 150–275 C and 400 cm1 is presented in Figure 2B. The FTIR spectra of
275–350 C, respectively. When the temperature was washed and acid-treated rice husk possess similar func-
raised from 360 to 650 C, losses in lignin content of the tional groups. In all the spectra, the intense broadband at
rice husk were observed. A higher temperature was 3434 cm1 represents an adsorbed molecule or O-H

FIGURE 2 (A) Thermogravimetric analysis (TGA) profile and (B) Fourier transform infrared (FTIR) spectra of MCM-41 and
Co/MCM-41
6 ARORA ET AL.

hydroxyl group, and the band at 2922 cm1 is ascribed to SEM, and results are shown in Figure 3. Figure 3A repre-
stretching vibrations of C-H2/C-H.[19] The characteristic sents the cobalt nanoparticles having a flower-like struc-
bands at 1733 cm1 and 1447 cm1 are designated to the ture with small nanorods of a size around 32.34 nm,
carbonyl group of C=O and –CH2 group, respec- which was found by taking an average of around
tively.[28,29] The band at 1059 cm1 represents symmetric 100 nanoparticles using ImageJ software.[20] The surface
stretching vibrations of the alkyl alcohol group. The pictograph of MCM-41 revealed that the material com-
small vibrations at ~802, ~673, and ~465 cm1 might be prised unevenly distributed spheroidal-shape particles
due to silicon atoms attached to the oxygen in the rice (Figure 3C).[37] However, cobalt nanoparticles-loaded
husk.[27,30] After calcination of rice husk, all the peaks MCM-41 (Figure 3E) have no significant difference in
related to organic compounds were lost due to the com- morphology, but a slightly thin film was observed on the
bustion of organic matter (mainly cellulose, hemicellu- support. The EDX results show the presence of Si and O
lose, and lignin). In the spectra of rice husk ash, a peak in a major portion of bare MCM-41, but cobalt-loaded
appears at 1632 cm1 owing to stretching vibration of the MCM-41 samples confirm Co species besides Si and
O-H group or absorbed water encapsulated in the struc- O. From the SEM and EDX results, it can be confirmed
ture of silica. The strong band at 1107 cm1 is ascribed to that cobalt species were successfully dispersed on
the structural siloxane (Si-O-Si) bond, representing the MCM-41.
formation of amorphous silica after calcination.[28] The nitrogen adsorption–desorption isotherms of
The vibrational modes near 808 and 467 cm1 which MCM-41 and Co/MCM-41 (5%, 400 C), and their
intensified after calcination is assigned to the formation corresponding pore size distributions are displayed in
of the Si-O bond.[31,32] Figure 4A,B. Both the samples exhibit typical type IV iso-
MCM-41 synthesized from rice husk ash exhibits a therms with the H1 hysteresis loop, which signifies the
strong band at 1096.96 cm1 associated with the asym- presence of cylindrical mesopores structure, presumably
metric stretching of the Si-O-Si bond, whereas the band due to the typical structure of MCM-41 according to
at 803 cm1 was attributed to symmetric and bending IUPAC nomenclature. The stability of liquid nitrogen
vibration of Si-O-Si.[6,28] Similar to rice husk ash, the meniscus in the material’s narrow pores may be the plau-
band at 467 cm1 in MCM-41 and cobalt-loaded MCM-41 sible cause of the formation of such a hysteresis loop.[25]
was assigned to the tetrahedral bending mode of Si-O.[25] Moreover, the pore structure of MCM-41 remained
The band at 1636.28 cm1 may be attributed to Si-OH undamaged even after cobalt nanoparticles were incorpo-
stretching vibration of a physically adsorbed water mole- rated into its framework. The pore size distribution of
cule.[18] In the IR spectra of all the samples of Co-loaded both samples illustrates the uniform distribution with the
MCM-41, the additional bands at 583.88 and 669.80 cm1 average pore size being in the range of 3–4 nm that corre-
appeared owing to the absorption band of Co3+-O sponds to the ordered structure of cylindrical pores
stretching mode and Co2+-O stretching vibrations of (Figure 4B).[38] As seen from the isotherm, monolayer
cobalt oxide nanoparticles in the framework of MCM- adsorption of nitrogen to the walls of MCM-41 and
41.[29] The peaks were barely visible when the loading cobalt-loaded MCM-41 takes place at a relative pressure
concentration of Co nanoparticles was less than 5 wt. below 0.25, mainly due to structural pore filling.
%, while the peak intensity increased with an increase Moreover, the multilayer adsorption on the particle’s
in loading amount to 10 wt.%, suggesting that the outer surface was evidenced in the relative pressure range
active metal was successfully incorporated on the sup- between 0.3–1.0. The isotherm is steeper in the relative
port material. pressure range of 0.8–1.0, which may be assigned to capil-
Purwaningsih et al.[33] investigated the rice husk ash lary condensation in inter-particulate secondary pores.[39]
composition using EDX analysis. They found a major The surface area and pore volume of MCM-41 decreased
portion of rice husk ash was silica and oxygen, and a slightly with the incorporation of active metal (Co/MCM-41)
minor portion was Na, Al, Fe, carbon, and so on. Acid from 532 to 490 m2/g and 0.506 to 0.453 cm3/g, respec-
treatment of rice husk effectively removes organic matter tively. Impregnation of cobalt nanoparticles on the
and other metal impurities. Kamari and Ghorbani[34] and MCM-41 framework may not cause any substantial
Bakar et al.[27] investigated the rice husk ash composition changes in average pore diameter, which decreased
using X-ray fluorescence (XRF). The acid treatment slightly from 3.653 to 3.627 nm.
removes metallic impurities from rice husk and signifi- Figure 5 depicts the transmission electron microscopy
cantly improves the purity of silica,[27,35,36] which serves (TEM) image of MCM-41 and Co/MCM-41 (5%, 400 C).
as a precursor for the synthesis of MCM-41. A highly ordered two-dimensional honey-type pore struc-
The morphology of Co3O4 nanoparticles, MCM-41, ture formed from series or parallel lines was observed in
and Co/MCM-41 (5%, 400 C) was investigated through the image (Figure 5A) when an electron beam was
ARORA ET AL. 7

F I G U R E 3 Scanning electron microscopy (SEM) and energy dispersive X-ray (EDX) analysis of (A) SEM image of Co3O4, (B) EDX of
Co3O4, (C) SEM image of MCM-41, (D) EDX of MCM-41, (E) SEM image of Co/MCM-41 (5%, 400 C), and (F) EDX of Co/MCM-41
(5%, 400 C)
8 ARORA ET AL.

FIGURE 4 (A) Nitrogen adsorption–desorption isotherm and (B) pore size distribution of MCM-41 and Co/MCM-41 (5%, 400 C)

F I G U R E 5 High-resolution image
of MCM-41 (A) along the pore and
(B) perpendicular to the pore;
Co/MCM41 (5%, 400 C) (C) along the
pore and (D) perpendicular to the pore
with the insert being the corresponding
particle

directed in the direction along with the pores of MCM-41 MCM-41 (Figure 5C). Some dark particles were observed in
framework. Furthermore, the TEM images were also taken Co/MCM-41, which may correspond to the dispersion of
by orienting the electron beam in the direction perpendicular cobalt nanoparticles on MCM-41. A clear honey-comb struc-
to the pores, which displays the channels and structure of ture was observed in the TEM image of Co/MCM-41
the MCM-41 (Figure 5B). From the micrographs, it can be (5%, 400 C), showing that MCM-41 retains its structure upon
believed that MCM-41 possesses a highly ordered hexagonal the loading of active metal into its framework elucidating
array of long-range silicate tubes.[25] Cobalt-loaded MCM-41 complete dispersion of cobalt nanoparticles in the structure
also showed a long-range hexagonal pore structure similar to of MCM-41 (Figure 5D).
ARORA ET AL. 9

F I G U R E 6 Full-range X-ray photoelectron spectroscopy (XPS) spectrum of (A) MCM-41, (B) full-range XPS spectrum of Co/MCM-41
(5%, 400 C), (C) C 1s, (D) O 1s, (E) Si 2p, and (F) Co 2p

The details of surface composition and chemical state MCM-41 was analyzed to confirm its interaction with sil-
linkage present in surface layers were studied in the ica and carbon. Curve fitting was performed by assuming
range 5–10 nm through the solid sample XPS spectra. the Gaussian-Lorentzian peak shape after subtracting
The XPS spectra of MCM-41 and Co/MCM-41 from the Shirley baseline. The C 1s XPS spectra of
(5%, 400 C) are depicted in Figure 6. The major photo- Co/MCM-41 (5%, 400 C) were deconvoluted into four
electron lines of C 1s, O 1s, and Si 2p were detected in major peaks centralized at the binding energy values of
both the spectra; also, Co species were identified in 284.06, 284.63, 286.01, and 287.73 eV, and these peaks
Co/MCM-41. The obtained findings are compatible with may be due to C-C, C=C, C-O, and C=O, respectively, as
the results of SEM–EDX. However, the deconvoluted shown in Figure 6C.[40,41] The appearance of the C bond
spectra of individual elements of MCM-41 are not shown in photoelectron spectra may be due to the surfactant
in Figure 6. The photoelectron spectra of cobalt-loaded decomposed in the MCM-41 framework. The spectra of O
10 ARORA ET AL.

F I G U R E 7 Optimization of reaction parameters to maximize yield and conversion: (A) influence of catalyst mass (molar ratio of DMC-
to-GLy: 4, temperature: 80 C, residence time: 2.5 h), (B) influence of molar mass of DMC-to-GLy (catalyst mass: 6 wt.%, temperature: 80 C,
residence time: 2.5 h), (C) influence of reaction temperature (catalyst mass: 6 wt.%, molar ratio of DMC-to-GLy: 3, residence time: 2.5 h),
(D) influence of residence time (catalyst mass: 6 wt.%, molar ratio of DMC-to-GLy: 3, temperature: 90 C), and (E) reusability study (catalyst
mass: 6 wt.%, molar mass of DMC-to-GLy: 3, reaction temperature: 90 C, residence time: 2 h)

1s were deconvoluted into three characteristic peaks: 531.13, different states of cobalt on MCM-41. Furthermore, the for-
532.20, and 532.69 eV; these deconvoluted peaks may be due mation of clear spin-orbit doublets located at around 781.08
to the presence of Co3O4, C-O, and Si-O-Si, respectively and 782.89 eV was due to the contribution of Co3+ associ-
(Figure 6D). The deconvoluted Si 2p spectra reveal the exis- ated to Co 2p3/2 oxidation states (Figure 6F), respectively.[44]
tence of peaks at a binding energy of 102.32, 103.00, and The shake-up satellite peaks positioned at 786.88 and
103.51 eV due to the presence of Si-C, SiO2, and Si-O, respec- 797.43 eV strongly correspond to Co2+ species. The occur-
tively (Figure 6E).[42,43] rence of different oxidation states might be due to the inter-
Apart from the silicate peaks, the XPS spectra of cobalt action of Co3+ with the pore walls of silicate followed by its
was deconvoluted into four peaks at binding energy, 781.08, reduction into Co2+ by -OH during calcination to remove
782.89, 786.88, and 797.43 eV, indicating the existence of surfactant.[45]
ARORA ET AL. 11

TABLE 1 Effect of basicity, yield, and conversion by varying calcination temperature and active metal loading

Basicity Basic Yield of Conversion


Catalyst (mmol/g) strength GLc (%) ±SE of GLy (%) ±SE TOF ±SE
MCM-41 5.0 H_ ≤ 7.2 4.8 0.5 5.0 0.4 0.3 0.0
Co/MCM-41(2%, 500 C) 7.1 7.2 ≤ H_ ≤9.8 19.1 1.8 19.2 2.2 1.2 0.1
Co/MCM-41(5%, 500 C) 8.2 7.2 ≤ H_ ≤9.8 40.2 2.2 42.0 1.7 2.8 0.2

Co/MCM-41(7%, 500 C) 9.2 7.2 ≤ H_ ≤9.8 32.5 1.9 35.8 2.9 2.3 0.1

Co/MCM-41(10%, 500 C) 9.8 9.8 ≤ H_ ≤15 16.6 0.5 17.2 0.6 1.1 0.1
Co/MCM-41(5%, 200 C) 5.8 H_ ≤7.8 11.3 1.5 11.5 1.8 0.7 0.0

Co/MCM-41(5%, 300 C) 6.5 7.2 ≤ H_ ≤9.8 25.1 3.3 25.7 2.9 1.7 0.2

Co/MCM-41(5%, 400 C) 8.7 7.2 ≤ H_ ≤9.8 60.2 2.6 63.8 1.9 4.2 0.0

Co/MCM-41(5%, 500 C) a
8.0 7.2 ≤ H_ ≤9.8 81.6 1.6 86.7 2.7 5.7 0.1

Note: Reaction condition for calcination temperature and active metal loading was conducted with a DMC-to-GLy molar ratio of 4, catalyst dose of 4 wt.%
(based on GLy mass), reaction temperature of 80 C, and reaction time of 2.5 h.
a
Regeneration study was conducted at optimized conditions: DMC-to-GLY molar ratio of 3, catalyst dosage of 6 wt.% (based on GLy mass), reaction
temperature of 90 C, and reaction time of 2 h.

During impregnation, cobalt nanoparticles coordinate understanding. TOF directly depends on the moles of
with the hydroxyl group. The aqueous cobalt complex GLc yield as well as the moles of GLy conversion.
interacts with the surface silanol group of support mate-
rial. The surface silanol group possesses amphoteric
character and can exist in the form of SiOH, SiO species, 3.2.1 | Effect of active metal loading and
the cationic exchange between the negatively charged sil- calcination temperature on GLc yield
ica and the positively charged cobalt ions is the
predominating process, resulting in the formation of The presence of basic sites on the catalyst surface
surface-bound Co3O4, Co2+, and Co-silicate like species enhances the transesterification reaction. It has been
with the support material.[46–48] The existence of these reported that these basic sites strongly depend on active
bonds is verified by XRD profile (Figure 1A), FTIR spec- metal loading as well as calcination temperature. Trans-
tra (Figure 2B), and XPS spectrum of cobalt (Figure 6F). esterification of GLy into GLc with the absence of cata-
lysts did not occur. In sequence, unloaded MCM-41
showed negligible conversion of GLy (~5% conversion),
3.2 | Catalyst activity and this may be due to a lack of basic sites.[49] In this
study, cobalt species (2–10 wt.%) on the MCM-41 frame-
Synthesized catalyst catalytic activity has been evaluated work showed better catalytic activity than bare MCM-41
to maximize the yield and selectivity of GLc. The proper- due to improvement of basic sites on the catalyst surface
ties of basicity and basic strength of catalyst highly (see Table 1). All the experiments were conducted thrice,
influenced the transesterification reaction. The batch and the standard error (SE) of each parameter was calcu-
transesterification results (Figure 7A–E) revealed that lated as shown in Figure 7. The GLy conversion and GLc
GLc selectivity had not fluctuated significantly, indicating yield increased from 19.2 ± 2.2% to 42.1 ± 1.7% and 19.1
that Co/MCM-41 has high selectivity for the conversion ± 1.8% to 40.2 ± 2.2%, respectively, with an increase in
of GLy into GLc without any by-product formation. This active metal cobalt from 2 to 5 wt.% on the MCM-41
may be due to the prepared Co/MCM-41 catalyst having framework. The positive trend may be due to the basic
moderate basic strength (7.2 < H_ < 15) and high basicity site concentration’s escalation from 7.1 to 8.7 mmol/g
(5–10 mmol/g). Das and Mohanty[2] reported that high (shown in Table 1). However, the GLy conversion, TOF,
selective GLc demands moderate basic strength and GLc yield, and selectivity showed a negative trend with a
strong basic sites of the catalyst. This might be the reason further increase in active metal cobalt from 7 to 10 wt.%,
why the catalyst developed in this study marked the high despite an increase in basic site concentration to
selectivity. In addition, the TOF has been calculated and 9.8 mmol/g. This may be due to multilayer dispersion of
presented (Figure 7A–D) for readers to obtain a thorough active metal at higher active metal loading on support,
12 ARORA ET AL.

which leads to decreased surface area by blockage/filling sites of the catalyst. With a further increase in calcination
of pores with an active metal, and the reactant species temperature from 400 to 500 C, the GLy conversion
may not access active sites to participate in the reaction. declined. At higher calcination temperature, sintering and
Calcination temperature strongly influences the surface agglomeration of particles on the active surface take place,
properties of metal oxides and the geometry of the support. which leads to the decline of basic sites.[2] From the TGA
The catalytic activity of 5 wt.% Co/MCM-41 catalyst cal- and XPS study, it was found that a trace amount of surface
cined at different temperatures from 200–500 C for 2.5 h directing agent was left over within the framework of
was used in the transesterification reaction for GLc synthe- MCM-41. The surface directing agent, CTAB, exhibits the
sis (Table 1). It was perceived that an increase in the calci- weak basicity dispersed inside the mesoporous channel,
nation temperature from 200 to 400 C, the basicity of the which is in favour of the transesterification reaction. Mori
catalyst, TOF, GLy conversion, GLc yield, and selectivity et al.[51] investigated the effect of residual templates with
showed a positive trend. This might be due to an improve- the MCM-41 framework. It was found that residual tem-
ment in basic site concentration from 5.8 to 8.7 mmol/g, plates (CTAB) within the MCM-41 exhibit superior cata-
which leads to higher catalytic activity on GLy conversion lytic activity than conventionally calcined catalyst
from 11.5 ± 1.8% to 63.8 ± 1.9%, respectively. However, at a without templates. This may be due to enhancement of
lower calcination temperature (200 C), the catalyst may basicity of the catalyst through surfactant and it is the
possess lower basic sites due to the incomplete decomposi- prime factor for the transesterification reaction.
tion of cobalt nitrate into oxides. But, at a higher calcination
temperature, cobalt nitrate converted into its oxide forms
(Co3O4), which were responsible for enhancing the basic 3.2.2 | Influence of reaction conditions
site, and thus, leading to higher catalytic activity for the
conversion of GLy. Parameswaram et al.[50] addressed that The yield of GLc and conversion of GLy were optimized
the number of oxide ions present on the catalyst surface upon studying the effect of catalyst dosage, molar ratio of
was responsible for an increase in the basicity and basic reactants, reaction time, and temperature.

F I G U R E 8 Comparative plots showing simulated and experimental concentration of GLy and GLc at (A) 70, (B) 80, (C) 90, and
(D) 100 C (DMC-to-GLy molar ratio of 3, Co/MCM-41 [5%, 400 C], catalyst dose of 6 wt.%)
ARORA ET AL. 13

TABLE 2 Equilibrium constants for the conversion of GLy to GLc at different reaction temperatures
Temperature k3 a ra  1004 Ea
( C) K1 ±SE K2 ±SE (min1) ±SE K4 ±SE K5 ±SE (min1) ±SE  1005 (kJ/mol) ±SE

70 2.6830 0.2341 2.8041 0.5623 0.0892 0.0045 2.6463 0.4559 2.6580 0.3652 1.3697 3.5635 58.2 2.4

80 2.7112 0.1653 2.8454 0.4896 0.1943 0.0395 2.6158 0.5236 2.5480 0.2362 1.0468 2.2895

90 2.7305 0.3631 2.9577 0.5231 0.4282 0.0651 2.5592 0.2651 2.4238 0.1542 0.2220 0.4865

100 2.7146 0.3562 2.9369 0.3581 0.4191 0.040 58 2.5471 0.3725 2.4138 0.2158 0.2350 0.1576

a
Reaction rate constant.

The effect of catalyst dose concerning GLy mass has GLy molar ratio of 3, and reaction time of 2.5 h on GLc
been investigated to enhance the reaction rate, keeping yield. The experimental results exhibited a gradual
other reaction parameters constant (i.e., the DMC to GLy improvement in GLc yield, TOF, and GLy conversion
molar ratio = 4, reaction temperature = 80 C, and the from 50.7 ± 2.2% to 91.3 ± 0.9%, 3.4 ± 0.4 to 6.5
reaction time = 2.5 h). Results are summarized in ± 0.4 h1, and 51.4 ± 1.2% to 98.4 ± 1.1%, respectively, as
Figure 7A, increasing the catalyst concentration from the reaction temperature increased from 70 to 90 C. This
1 to 6 wt.% (concerning the GLy mass), the GLy conver- might be because, at higher temperatures, the kinetic
sion, GLc yield, and TOF were increased from 30.5 energy of the liquid mixture increases, resulting in a
± 0.7% to 75.8 ± 1.5%, 29.9 ± 1.2% to 71.3 ± 0.5%, and decrease in the reactant molecules’ viscosity, which leads
2.0 ± 0.1 to 5.0 ± 0.2 h1, respectively. The positive trend to the improvement in miscibility, thus enhancing the
might be attributed to the increase in active/basic sites yield and conversion.[53] However, a decrease in the
with an increase in catalyst mass, which maximizes the selectivity of GLc was observed with a further increase of
GLc yield.[52] A decline in yield of GLc was witnessed temperature from 90 to 100 C. The decline of yield could
when the catalyst concentration increased from 6 to 8 wt. be attributed to the decomposition of the formed mole-
% in the reaction mixture. This might be because the cule into other by-products due to opening of the GLc
agglomeration of the particles at higher catalyst loading ring on excess available basic sites on the surface of the
hinders the accessibility of the reactants to the active sites catalyst.[7]
of catalyst, and thereby stabilizes the conversion of GLy. The effect of reaction time was investigated to maxi-
Algoufi et al.[52] reported that increasing catalyst mass mize the GLc yield with a set of experiments conducted
beyond a threshold value would induce sufficient at the optimized conditions at a regular time interval of
mass transfer resistance. 0.5–2.5 h (Figure 7D). Both the GLy conversion and GLc
The influence of reactant mass (the DMC-to-GLy yield increased with an increase in reaction time. In the
molar ratio) was investigated in the range of 1–7 under reaction time of 0.5 h, 63.3 ± 2.5% GLc yield and 64.5
constant reaction conditions (Reaction time = 2.5 h, reac- ± 1.5% GLy conversion were observed. The GLc yield
tion temperature = 80 C, and catalyst dose = 6 wt.%) and GLy conversion rose to 94.1 ± 0.7% and 98.7
(Figure 7B). The GLc yield, GLy conversion, and TOF ± 1.2%, respectively, when the reaction time was pre-
showed a positive trend from 57.9 ± 1.1% to 78.7 ± 2.6%, ceded to 2 h. The minor increase in GLy conversion
58.6 ± 2.9% to 80.9 ± 1.5%, and 3.9 ± 0.09 to 4.5 and a slight decrease in GLc yield were observed with
± 0.2 h1, respectively, with an increase of DMC-to-GLy an increase in the reaction time from 2 to 2.5 h, owing
molar ratio from 1 to 3. This may be due to a shift of to decarboxylation of GLc into glycidol on the active
equilibrium towards the forward direction when increas- sites of the catalyst upon prolonging the reaction
ing the reactant molar ratios. However, a further increase time.[53] Though glycidol quantification was moni-
in the DMC-to-GLy molar ratio from 3 to 7 would reduce tored, glycidol was not identified. This may be because
the conversion, yield, and TOF because dilution of cata- of the low concentration of glycidol present in the reac-
lyst concentration in the reaction mixture occurred when tion mixture. Hence, the reaction time of 2 h was taken
the DMC was added in a large amount. The observed as an optimized time to conduct the transesterification
trends are in good agreement with the studies on catalyst of GLy into GLc.
mass effect.[9]
The experiments have been conducted at different
temperatures to investigate the critical importance of 3.3 | Reusability and catalyst
reaction temperature in the transesterification reaction stability test
for the synthesis of GLc. Figure 7C demonstrates the
influence of reaction temperature ranging from 70– The feasibility of heterogeneous catalysts in an industrial
100 C, with a constant catalyst dose of 6 wt.%, a DMC-to- application needs to meet the reusability criteria without
14 ARORA ET AL.

rapid deactivation. By considering this, the reusability of chemical reaction kinetics. The catalytic transesterification
Co/MCM-41 was investigated via reuse experiments reaction of GLy with DMC follows three significant steps:
(Figure 7E). The method implemented to reuse the spent (i) adsorption of reactants (GLy and DMC), (ii) the surface
catalyst was the same as that in Arora et al.[21] without reaction between GLy and DMC molecules on the active
adopting any regeneration technique. Spent catalyst was site, and (iii) desorption of products and unconverted reac-
washed with methanol, dried, and reused in the subsequent tants from the active site (GLc and methanol). To describe
reaction. Results concluded that cobalt impregnated MCM- the possible dual site catalytic process, the Langmuir-Hin-
41 could be used for four cycles without sustainable loss in shelwood-Hougen-Watson (LHHW) model was used.[55]
its activity. After the fourth cycle of reusability, the GLc yield The kinetic model Equation (8) was adopted as developed
and GLy conversion declined slightly from 94.1 ± 0.7% to by Arora et al.[21] for GLy and DMC transesterification. The
~81.6 ± 1.6% and 98.7 ± 1.2% to 86.7 ± 2.7%, respectively. elucidated derivation of Equation (8) is provided in the
This declining trend may be due to loss of basic sites and Supporting Information (SI, Appendix S1):
leaching of active metal (last line of Table 1). After the fourth
cycle, the spent catalysts were separated, dried, and then dCGLy k0 CGLy
¼ 2 ;
used to investigate their surface chemistry and crystallinity dt
1 þ KGLy CGLy þ KGLy CGLy þ KCGLc þ KCMM ð8Þ
using FTIR and XRD. It was seen from the spectra that the GLc

intensity of the peak of active metal decreased in XRD and k0 ¼ kSR KGLy KDMC CDMC C2t
FTIR profiles of the spent catalysts, which might be responsi-
ble for the loss in activity of the catalyst.[54]
The parity plot between the experimental data and simu-
lated data of GLc and GLy is shown (Figure S1A,B in SI,
3.4 | Green parameters Appendix S1), and an excellent fit with an error of 20%
and 10%, respectively, was witnessed. Equation (8) was
The present study’s greener approach and eco-friendly solved to compute different rate constants of the reaction
nature were quantitatively accessed by the E-factor and using stiff integrators-ODE 15s from MATLAB. All the
process mass intensity using Equations (6) and (7). The parameters were taken in the range of positive real num-
lower the E-factor, the lower the waste generated during bers. For each run, the value of the concentration of each
the process. Thereby the process is considered eco- species was passed to the objective function, Equation (9),
friendly for the environment: and the best value was selected based on the value pro-
vided by the ‘Fmincon’ function.
Total waste ðgÞ The concentrations of GLy and GLc (simulated and
E  factor ¼ ð6Þ
Product ðgÞ experimental) were compared. The graphs are shown in
Figure 8A–D. From the results, it can be concluded that
where total waste = total mass of material used in the the model shows an excellent fit between the experimen-
process  total mass of product formed: tal and simulated data of GLy and GLc. The criteria for
the determination of rate constant were that the value of
Total mass used in the process ðgÞ the sum of squares (SSE), root mean square error
PMI ¼ ð7Þ
Mass of final product ðgÞ (RMSE), and chi-square (χ2) should tend to 0, while the
coefficient of determination (R2) should be close to
Since methanol is the only major by-product of trans- 1. The mathematical formula used for the calculation of
esterification of GLy into GLc, which could be further SSE, RMSE, χ2, and R2 is provided in the SI[56]:
used in the biodiesel production plant, the recoverability
of solvent and reusability of catalyst were not considered N 
X 2  2 
while calculating the values of these parameters. E-fac- f¼ CiGLy,exp  CiGLy,pred þ CiGLc,exp  CiGLc,pred
i¼1
tor = 1.1 and PMI = 2.1 were calculated for the produc-
ð9Þ
tion of GLc, showing the minimum waste generated
during the process.
The values of statistical parameters are presented in the
SI, which shows that SSE, RMSE, and χ2 tend to nearly
3.5 | Reaction kinetics zero. In contrast, the higher values of the R2 approxi-
mately incline to 1, illustrating a good fit with the model.
In order to get an in-depth understanding of heteroge- The values of equilibrium constants at temperatures
neous catalytic reactions, it is essential to evaluate its between 70–100 C is shown in Table 2. When the
ARORA ET AL. 15

reaction was conducted at higher temperatures, the reac- Verraboina Subbaramaiah https://orcid.org/0000-0003-
tion rate increased due to the movement of reaction 4457-6725
equilibrium towards a forward direction. It can be easily
perceived from the table that the apparent reaction rate RE FER EN CES
constant increases with the rise in reaction temperature [1] Y. T. Algoufi, G. Kabir, B. H. Hameed, J. Taiwan Inst.
from 70 to 90 C, signifying that the concentration of Chem. E. 2017, 70, 179.
GLy and GLc attains equilibrium rapidly with the [2] B. Das, K. Mohanty, Journal of Environmental Chemical
increase in reaction temperature. The transesterification Engineering 2019, 7, 102888.
[3] M. A. Do Nascimento, L. E. Gotardo, R. A. C. Le~ao, A. M. De
reaction activation energy was calculated from the
Castro, R. O. M. A. De Souza, I. Itabaiana, ACS Omega 2019,
Arrhenius plot and found to be 58.2 ± 2.4 kJ/mol. 4, 860.
[4] P. U. Okoye, A. Z. Abdullah, B. H. Hameed, J. Taiwan Inst.
Chem. E. 2016, 68, 51.
4 | C ON C L U S I ON S [5] S. Arora, V. Gosu, U. K. A. Kumar, V. Subbaramaiah, Int.
J. Chem. React. Eng. 2020, 18, 1.
The present study successfully utilized agricultural waste [6] P. Sikarwar, U. K. A. Kumar, V. Gosu, V. Subbaramaiah,
(rice husk) to prepare a mesoporous structure catalyst for Journal of Environmental Chemical Engineering 2018, 6, 1736.
[7] P. Kumar, P. With, V. C. Srivastava, R. Gläser, I. M. Mishra,
the valorization of GLy obtained from the biodiesel
Ind. Eng. Chem. Res. 2015, 54, 12543.
industry. Synthesized catalyst portrays well-defined hex- [8] M. Malyaadri, K. Jagadeeswaraiah, P. S. Sai Prasad, N.
agonal structure exhibiting high thermal stability up to Lingaiah, Appl. Catal. A-Gen. 2011, 401, 153.
800 C. The surface area and pore volume of Co/MCM-41 [9] X. Song, Y. Wu, F. Cai, D. Pan, G. Xiao, Appl. Catal. A-Gen.
(5%, 400 C) declined compared with bare MCM-41 from 2017, 532, 77.
532 to 490 m2/g and 0.506 to 0.453 cm3/g, respectively. [10] G. Parameswaram, P. S. N. Rao, A. Srivani, G. N. Rao, N.
Basic sites on the catalyst surface significantly influenced Lingaiah, Molecular Catalysis 2018, 451, 135.
[11] P. Liu, M. Derchi, E. J. M. Hensen, Appl. Catal. A-Gen. 2013,
the catalyst’s activity towards GLc yield. The catalyst
467, 124.
developed sustained its activities for transesterification
[12] D. Li, H. Min, X. Jiang, X. Ran, L. Zou, J. Fan, J. Colloid Interf.
reactions up to the fourth cycle without any regenera- Sci. 2013, 404, 42.
tion. The value of K1, K2, K3, K4, and kˈ in the devel- [13] M. J. Hülsey, H. Yang, N. Yan, ACS Sustain. Chem. Eng. 2018,
oped differential equation was solved through 6, 5694.
MATLAB’s ODE 15s tool. The activation energy was [14] A. Gidde, M. R. Jivani, in Proc. of the Int. Conf. on Cleaner Tech-
found to be 58.2 ± 2.4 kJ/mol. The obtained E-factor nologies and Environmental Management, PEC, Pondicherry,
value shows that the process is environmentally India 2007, p. 586.
[15] F. W. Chang, M. S. Kuo, M. T. Tsay, M. C. Hsieh, Appl. Catal.
friendly and green.
A-Gen. 2003, 247, 309.
[16] H. B. Dizaji, T. Zeng, I. Hartmann, D. Enke, T. Schliermann,
ACK NO WLE DGE MEN TS V. Lenz, M. Bidabadi, Appl. Sci. 2019, 9, 1.
This work was supported by Department of Science and [17] N. Sapawe, N. Surayah Osman, M. Zulkhairi Zakaria, S.
Technology, Government of India through Early Career Amirul Shahab Syed Mohamad Fikry, M. Amir Mat Aris,
Research grant (ECR/2015/000085, dated 20 August Mater. Today-Proc. 2018, 5, 21861.
2016). Ms. Shivali Arora would like to thank Malaviya [18] H. M. A. Hassan, M. A. Betiha, R. F. M. Elshaarawy, M. Samy
National Institute of Technology, and the Department of El-Shall, Appl. Surf. Sci. 2017, 402, 99.
[19] M. Bhagiyalakshmi, L. J. Yun, R. Anuradha, H. T. Jang,
Higher Education, Ministry of Education, Government of
J. Porous Mat. 2010, 17, 475.
India for the award of a fellowship. [20] M. Salavati-Niasari, F. Davar, M. Mazaheri, M. Shaterian,
J. Magn. Magn. Mater. 2008, 320, 575.
P EE R R EV IE W [21] S. Arora, V. Gosu, V. Subbaramaiah, Molecular Catalysis 2020,
The peer review history for this article is available at 496, 111188.
https://publons.com/publon/10.1002/cjce.24284. [22] Z. Zhu, G. Lu, Z. Zhang, Y. Guo, Y. Guo, Y. Wang, ACS Catal.
2013, 3, 1154.
[23] R. V. Krishnarao, M. M. Godkhindi, Ceram. Int. 1992, 18, 243.
DATA AVAILABILITY STATEMENT
[24] R. V. Krishnarao, M. M. Godkhindi, P. G. I. Mukunda, M.
All data included in this study are available upon request
Chakraborty, J. Am. Ceram. Soc. 1991, 74(11), 2869.
by contact with the corresponding author. [25] C. Huo, J. Ouyang, H. Yang, Sci. Rep. 2015, 4, 1.
[26] W. Roschat, T. Siritanon, B. Yoosuk, V. Promarak, Energ.
ORCID Convers. Manage. 2016, 119, 453.
Vijayalakshmi Gosu https://orcid.org/0000-0001-6510- [27] R. A. Bakar, R. Yahya, S. N. Gan, Procedia Chem. 2016,
9873 19, 189.
16 ARORA ET AL.

[28] S. Arora, V. Gosu, U. K. A. Kumar, V. Subbaramaiah, J. Clean.  Szegedi, M. Popova, C. Minchev, J. Mater. Sci. 2009, 44,
[46] A.
Prod. 2021, 295, 126437. 6710.
[29] K. Tang, X. Hong, Energ. Fuel. 2016, 30, 4619. [47] E. Van Steen, G. S. Sewell, R. A. Makhothe, C. Micklethwaite, H.
[30] T. H. Liou, S. J. Wu, J. Hazard. Mater. 2009, 171, 693. Manstein, M. De Lange, C. T. O’Connor, J. Catal. 1996, 162, 220.
[31] D. An, Y. Guo, Y. Zhu, Z. Wang, Chem. Eng. J. 2010, 162, 509. [48] S. Madadi, J. Y. Bergeron, S. Kaliaguine, Catal. Sci. Technol.
[32] F. Adam, K. Kandasamy, S. Balakrishnan, J. Colloid Interf. Sci. 2021, 11, 594.
2006, 304, 137. [49] K. Lanjekar, V. K. Rathod, Journal of Environmental Chemical
[33] H. Purwaningsih, S. Raharjo, V. M. Pratiwi, D. Susanti, A. Engineering 2013, 1, 1231.
Purniawan, Mater. Sci. Forum 2019, 964, 88. [50] G. Parameswaram, M. Srinivas, B. Hari Babu, P. S. Sai Prasad,
[34] S. Kamari, F. Ghorbani, Biomass Conversion and Biorefinery N. Lingaiah, Catal. Sci. Technol. 2013, 3, 3242.
2020, https://doi.org/10.1007/s13399-020-00637-w. [51] H. Mori, K. Takuya, Y. Ikurumi, S. Yamashita, Chem.
[35] T. Sirisoontornpanit, A. Wongkoblap, S. Junpirom, Key Eng. Commun. 2013, 49, 10468.
Mater. 2017, 729, 24. [52] Y. T. Algoufi, U. G. Akpan, G. Kabir, M. Asif, B. H. Hameed,
[36] K. Kordatos, A. Ntziouni, L. Iliadis, V. Kasselouri-Rigopoulou, Energ. Convers. Manage. 2017, 138, 183.
J. Mater. Cycles Waste 2013, 15, 571. [53] Y. T. Algoufi, B. H. Hameed, Fuel Process. Technol. 2014, 126, 5.
[37] B. M. Abu-Zied, M. M. Alam, A. M. Asiri, W. Schwieger, [54] S. Wang, J. Wang, P. Sun, L. Xu, P. U. Okoye, S. Li, L. Zhang,
M. M. Rahman, Colloid. Surface. A 2019, 562, 161. A. Guo, J. Zhang, A. Zhang, J. Clean. Prod. 2019, 211, 330.
[38] G. V. Mamontov, A. S. Gorbunova, E. V. Vyshegorodtseva, V. I. [55] P. Devi, U. Das, A. K. Dalai, Chem. Eng. J. 2018, 346, 477.
Zaikovskii, O. V. Vodyankina, Catal. Today 2019, 333, 245. [56] H. A. Hamid, R. Yunus, U. Rashid, T. S. Y. Choong, A. H.
[39] V. Subbaramaiah, V. C. Srivastava, I. D. Mall, Ind. Eng. Chem. Al-Muhtaseb, Chem. Eng. J. 2012, 200-202, 532.
Res. 2013, 52, 9021.
[40] Y. Li, J. Liu, D. He, Appl. Catal. A-Gen. 2018, 564, 234.
[41] P. Sikarwar, U. K. A. Kumar, V. Gosu, V. Subbaramaiah, SU PP O R TI N G I N F O RMA TI O N
J. Chem. Eng. Data 2018, 63, 2975. Additional supporting information may be found online
[42] R. Gostynski, R. Fraser, M. Landman, E. Erasmus, J. in the Supporting Information section at the end of this
Conradie, J. Organomet. Chem. 2017, 836–837, 62.
article.
[43] F. J. Grunthaner, P. J. Grunthaner, R. P. Vasquez, B. F. Lewis,
J. Maserjian, J. Vac. Sci. Technol. 1979, 16, 1443.
[44] B. Solsona, E. Ayl on, R. Murillo, A. M. Mastral, A. Monzonís, How to cite this article: S. Arora, V. Gosu,
S. Agouram, T. E. Davies, S. H. Taylor, T. Garcia, J. Hazard.
V. Subbaramaiah, T. C. Zhang, Can. J. Chem. Eng.
Mater. 2011, 187, 544.
[45] S. Shen, J. Chen, R. T. Koodali, Y. Hu, Q. Xiao, J. Zhou, X.
2021, 1. https://doi.org/10.1002/cjce.24284
Wang, L. Guo, Appl. Catal. B-Environ. 2014, 150–151, 138.

You might also like