Next Article in Journal
Battery Dynamic Balancing Method Based on Calculation of Cell Voltage Reference Value
Previous Article in Journal
An Ab Initio RRKM-Based Master Equation Study for Kinetics of OH-Initiated Oxidation of 2-Methyltetrahydrofuran and Its Implications in Kinetic Modeling
Previous Article in Special Issue
Biological Hydrogen Production from Biowaste Using Dark Fermentation, Storage and Transportation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermocatalytic Decomposition of Sesame Waste Biomass over Ni-Co-Doped MCM-41: Kinetics and Physicochemical Properties of the Bio-Oil

1
National Centre of Excellence in Physical Chemistry, University of Peshawar, Peshawar 25120, Pakistan
2
Department of Chemistry, Quaid-i-Azam University, Islamabad 45320, Pakistan
3
School of Chemistry, University of the Punjab, New Campus, Lahore 54590, Pakistan
4
Department of Chemistry, GC Peshawar, University of Peshawar, Peshawar 25120, Pakistan
*
Author to whom correspondence should be addressed.
Energies 2023, 16(9), 3731; https://doi.org/10.3390/en16093731
Submission received: 26 March 2023 / Revised: 15 April 2023 / Accepted: 24 April 2023 / Published: 27 April 2023
(This article belongs to the Special Issue New Challenges in Waste Biomass)

Abstract

:
The increase in industrialization and development has tremendously diminished fossil fuel resources. Moreover, the excessive use of fossil fuels has resulted in the release of various toxic gases and an increase in global warming. Hence, necessitating the need to search for a renewable energy source. In this study, sesame waste biomass (SWB), which is abundantly available in Pakistan, has been used as feedstock for obtaining bio-oil using the pyrolysis technique. Pyrolysis was carried out using thermogravimetry and a pyrolysis chamber. Firstly, thermogravimetric analysis was performed on biomass with/without a laboratory synthesized catalyst Ni/Co/MCM-41 in nitrogen at different temperature programmed rates of 5, 10, 15, and 20 °C/min. A four-stage weight loss was observed that pointed toward the vaporization of water, and degradation of hemicelluloses, cellulose, and lignin. The kinetics parameters were determined using the Kissinger equation. The activation energy for the decomposition reaction of hemicelluloses, cellulose, and lignin, without catalyst, was observed as 133.02, 141.33, and 191.22 kJ/mol, respectively, however, with catalyst it was found as 91.45, 99.76, and 149.65 kJ/mol, respectively. In the catalyzed reaction the results showed the lowest activation energy, which is an indication of the fact that the catalyst is successful in reducing the activation energy to a sufficient level. As the TG/DTG showed active degradation between 200 and 400 °C, therefore, the waste sesame biomass over Ni-Co/MCM-41 was pyrolyzed within the same temperature range in the pyrolysis chamber. Temperature and time were optimized for maximum oil yield. A maximum oil yield of 38% was achieved at 330 °C and 20 min. The oil obtained was studied using GCMS. The physicochemical characteristics of the oil were assessed, and it was found that if the oil was upgraded properly, it could serve as a fuel for commercial use.

1. Introduction

Energy is a critical economic factor in the development of a country [1]. Fossil fuels have been the major sources of energy for many of the past decades; however, with a sudden rise in industries, the fossil fuel reserves are continuously depleting [2]. Moreover, the use of fossil fuels results in the emission of different toxic gases into the environment, which also contributes to the greenhouse effect. Hence, it necessitates the need to search for renewable energy [3]. Therefore, scientists across the world are looking for renewable energy sources that are not exhaustible and environmentally friendly [4]. Thus, more attention is being paid to biomass due to its low cost and eco-friendly nature [5,6,7].
Energy is obtained from biomass using different techniques, i.e., combustion, biochemical, and pyrolysis. In combustion, the biomass is fully oxidized at elevated temperatures and the heat produced is used instantaneously for heating or generating power. This process is not preferred due to the drawbacks associated with it, i.e., low efficiency, unwanted ash, and increased carbon dioxide [3]. The two commonly used biochemical processes for biomass conversion into energy are anaerobic digestion and fermentation; however, these processes are more time-consuming and less efficient. Therefore, pyrolysis is one of the best options as it produces liquid fuel, non-condensable gases, and solid fuel at modest temperatures [8,9].
Although many pyrolysis techniques are being in-vogue, the pyrolysis of biomass is usually carried out by slow pyrolysis, which is quite favorable for the production of olefins and aromatics [10]. The quality and quantity of products that are obtained are greatly dependent on the composition of the biomass, reaction time, temperature, and type of catalyst [11,12]. As oxygenated compounds are found in abundance in bio-oil, they cause it to decrease in stability, heating value, volatility, and pH, while increasing its viscosity. Hence, to improve the value of fuel, a catalyst needs to be used in the pyrolysis chamber. The catalyst initiates a series of reactions resulting in the conversion of big molecules to light hydrocarbons [13,14].
The most widely used catalysts for the pyrolysis of biomass are zeolites and metal oxides [15,16,17]. Natural materials have also been used as catalysts for such purposes [18]. Any porous material, whether natural or manmade, can be used for this purpose due to the pore volume and an accessible high surface area. In this connection, MCM-41 has a framework of cylindrical mesospores arranged in one dimension, with a large surface area, and a pore volume that has been widely utilized by many researchers.
Adam et al. [19] investigated the degradation of spruce wood over Al-MCM-41. The products obtained were analyzed by GCMS, while the acetic acid, furfural, and furans were detected as the major hydrocarbons in the oil. Nilsen et al. [20] incorporated Zn, Cu, and Fe into Al-MCM-41 and the modified catalysts improved the quality of oil; however, Zn–Al-MCM-41 was observed to more effectively reduce the coke production. Antonakou et al. [21] concluded that a low ratio of Si/Al in Al/MCM-41 had a positive effect on the bio-oil yield. Fontes et al. [22] studied the pyrolysis of biomass using Ti-MCM-41 with Si/Ti in different ratios. The Ti-MCM-41 with a Si/Ti ratio of 50 was found to be more efficient in yielding a high quantity of hydrocarbons. Li et al. [23] used loaded MCM-41 with 1, 5, and 10% of La in rape straw pyrolysis. The 5% La-loaded catalyst showed good efficiency with a hydrocarbon content of 35% and was also successful in lowering the coke content on the catalyst surface, resulting in an increase in the selectivity for aromatic hydrocarbons.
Mesoporous Al/MCM-41 has also been used in combination with microporous ZSM-5 for biomass pyrolysis and MCM-41 was observed to obstruct coke development on ZSM-5, while both helped each other in improving the yield [24,25]. Ratnasari et al. [26] studied the kinetics of biomass pyrolysis over a mixture of H-ZSM-5 and Al/MCM-41 and noted a decrease in the activation energy of the reaction in the presence of a catalyst. Ni/P/MCM-41 was also tried by Li et al. [27] for levoglucosenone production from biomass and revealed its best selectivity and activity in the production of levoglucosenone.
From the aforementioned literature, we concluded that the doping of another material over MCM-41 could further improve the catalytic behavior. For this purpose, different metals, such as Zn, Al, Cu, and Fe were used to increase the catalytic efficiency of MCM-41, yet an extensive gap still exists in the discovery of the influence of more metals on the efficiency, selectivity, and activity of MCM-41. Therefore, a 9% Ni/8% Co/MCM-41 composition, which has not been explored previously, was attempted for the pyrolysis of biomass waste. Among the biomass, sesame (botanical name: Sesamum indicum L. and a local name of Til) is the most significant plant grown abundantly in Pakistan. The utilization of sesame timber in Pakistan is restricted only to cooking food and baking bricks, whereas most of the timber is dumped and wasted. Therefore, the present work focused on the use of Ni/Co/MCM-41 as a catalyst for bio-oil yield from sesame biomass and to improve the quality of the oil being produced. Moreover, the effects of the catalyst on the reduction of the maximum degradation temperature and activation energy were also evaluated.

2. Experimental

2.1. Materials and Method

Sesame plant waste was gathered from a harvested crop in the district Multan, south of Punjab, Pakistan. These were crushed into minute pieces and sieved through mesh no. 35 to provide suitable particles for pyrolysis. Then, it was washed, dried, and mixed with the required amount of catalyst for further studies. Chemicals, i.e., ammonia solution 32%, phosphoric acid, tetraethyl orthosilicate, cetyltrimethylammonium bromide, cobalt nitrate, nickel nitrate, and citric acid were procured from Sigma-Aldrich and Sharlu. N2 (99.99% pure) was purchased from New Khyber Oxygen Limited Peshawar.

2.2. Synthesis and Characterization of Catalyst

Ni/Co/MCM-41 was synthesized by templating the mechanism by using the support materials of MCM-41 and pure cetyltrimethylammonium bromide, tetraethyl orthosilicate, ammonia, and water. Firstly, MCM-41 was prepared by dissolving 1 g of cetyltrimethylammonium bromide in 102.5 mL of ammonium hydroxide and 135 mL water, before the solution was stirred for 35 min to produce a rod-style homogeneous micelle. Then, 5 mL of tetraethyl orthosilicate was poured into it, and it was stirred for 2 h until a snowy slurry was attained. This slurry was repeatedly washed with water, dried, and calcined at 550 °C for 6 h.
Secondly, 9% Ni/8% Co/MCM-41 was synthesized using the sol-gel method, utilizing the prepared MCM-41 as support. To synthesize the 9% Ni/8% Co/MCM-41, 1.1 g of citric acid was dissolved in 35 mL of water before 0.535 g nickel nitrate hexahydrate and 0.476 g cobalt nitrate hexahydrate were added into it. Finally, 1 g of MCM-41 was dissolved into the solution, and it was stirred at 60 °C unless a gel appeared. The formed gel was aged for 3 days, without stirring, at ambient temperature, and then, calcined at 550 °C for 6 h and stored in a container for further study.
The morphology of the prepared Ni/Co/MCM-41 was assessed using a scanning electron microscope JSM5910 JEOL, Akishima-shi, Japan. The crystalline/amorphous feature was observed through XRD (JDX-3532, JEOL Japan). The elemental analysis was conducted using EDX (INCA100, Oxford instruments, Oxford, UK). The surface features were determined through NOVA 2200e Quantachrome US, surface area, and a pore size analyzer.

2.3. Preparation of Biomass for Analysis

Three separate samples raw, acid-washed, and mixed with a catalyst were prepared for pyrolysis. The demineralized biomass was prepared by treating it with a 2 N solution of phosphoric acid, filtering it through a vacuum pump, and washing and drying it at 105 °C for 20 h. The catalyst-loaded biomass was prepared by impregnating it with 5% of 9% Ni/8% Co/MCM-41.

2.4. Thermogravimetry

Sesame waste biomass, in the absence/presence of the catalyst, was pyrolyzed in a thermobalance (Q500 TG coupled with DSC 8000) at various temperature programmed rates and the resulting information was applied to determine the kinetic parameters, using the Kissinger equation (Equation (1)):
ln ( β T m 2 ) = ln [ A . R E a ] E a R T m
where β and Tm stand for heating rate and peak degradation temperature, other well-known terms have their respective meanings. Ea and A-factor were determined from the plots of ln( β / T m 2 ) versus 1/Tm [28].

2.5. Pyrolysis of Biomass

The degradation of the biomass was performed in a salt bath from 300 to 400 °C in nitrogen. The temperature of the pyrolysis chamber was monitored by a temperature controller (Shimaden, E234017, Tokyo, Japan), via thermocouple. Experiments were performed in a glass vessel attached to a condenser and the oil obtained was studied through GCMS (DSQ II, Thermo Scientific, Waltham, MA, USA). The data obtained were matched with an NIST MS library and the peaks were identified accordingly.

2.6. Physicochemical Properties

The physicochemical properties, such as density, viscosity, fluidity, and the specific gravity of the oil obtained were determined. For density, the Pycnometer, with a capacity of 25 mL, was used, while the viscosity was determined using the Ostwald viscometer, whereby water was used as the standard liquid. Fluidity was calculated by taking the reciprocal of the determined bio-oil viscosity and for specific gravity determination, the ratio of the density of the bio-oil to the density of the water was noted.

3. Results and Discussion

3.1. Characterization of Catalyst

Figure 1a,b show the SEM pictures of the MCM-41 and 9% Ni/8% Co/MCM-41, respectively. The micrographs show spongy features outlining the spherical and the ordered morphology with less agglomeration, which shows that the surface features of the MCM-41 have not been changed following the addition of Ni and Co because they are less than 10%, which was also confirmed by using XRD and with the appearance of an intense peak. These results are in accordance with the findings by Grün et al. [29], who noted that the as-prepared MCM-41 particles were spherical in shape and less agglomerated. Likewise, Nejat et al. [30] assessed the morphologies of MCM-41, Ni/MCM-41, and Co/MCM-41 using a field emission scanning electron microscopy and observed that the particles were spherical in shape and uniform in size with a diameter of 50 nm and that there was no change in their morphology following the incorporation of Co and Ni. Wu et al. [31] observed no change in the surface morphology of Ni-MCM-41 with Ni loadings of 5, 10, and 20%, however, at 40%, a prominent change in the surface features was observed, with an increase in the bulky NiO outside the pores.
The chemical composition of the catalyst was determined through EDX. Figure 1c shows no traces of nickel or cobalt in the pure MCM-41, whereas, in Figure 1d, the prominent peaks for Ni and Co can be seen to show a successful incorporation of Ni and Co onto MCM-41. Moreover, the presence of Si and O, in both figures, points toward the silicate nature of MCM-41, which has also been previously confirmed by Parangi et al. [32]. Parangi and co-authors, while performing EDX analysis of 12TPA-MCM-41, observed that Si and O had an atomic percentage of 30.44 and 67.40, respectively. Abrokwah et al. [33] observed from the EDX of a different metal-impregnated MCM-41 that Cu, Pd, Zn, and Sn were highly distributed in MCM-41, whereas Ni and Co showed little dispersion with only a slight agglomeration.
Figure 1e shows the XRD of MCM-41. The broad peak at 2θ = 22.7 is due to siliceous MCM-41 [33]. These observations are consistent with some earlier studies. Salam et al. [34] performed XRD on sulfated Zr-MCM-41 and observed a strong peak at 2θ = 21°. However, in the case of 9% Ni/8% Co/MCM-41, some broad lines were noted that point toward the amorphous nature of the catalyst. Moreover, in addition to the broad peak at 2θ = 23.2°, some weak signs were also observed at 2θ = 36.3°, 43.1°, and 60.6° due to the Ni and Co incorporation onto MCM-41, as shown by Figure 1f [35,36]. The crystallite size determined from the XRD data was noted to be 32 nm. The investigated data are in consonant with the findings reported by Nejat et al. [30]. The absence of an intense peak may be due to a decrease in the amount of the doped materials over the support, i.e., 9% Ni and 8% Co, respectively. The outcomes are in agreement with the reported observation by Wu et al. [31]. The authors impregnated MCM-41 with 5,10, 20, and 40% Ni and analyzed the samples using XRD. In all the samples, a broad peak was noted at 23°, yet, no peak was observed for the 5% and 10% loaded samples. For the 20% loaded sample, the obtained peaks were also not prominent, while, three sharp peaks were observed at 37°, 43°, and 64° in the 40% loaded sample. Liu et al. [37] attributed the broadness of the XRD peak of Ni/Zr/MCM-41 to the impregnation of a second metal ion, resulting in the substitution of Si+4 and a change in the T–O–T bond angle and a decrease in the order of mesostructures. The specific surface area of MCM-41 and 9% Ni/8% Co/MCM-41 was assessed using the Brunauer–Emmett–Teller method and the pore size was assessed using the Barrett–Joyner–Halenda method. MCM-41 was observed to possess a surface area of 511.91 m2/g, a pore volume of 0.017 cm3/g, and a pore radius of 15.16 A°, whereas the same properties for 9% Ni/8% Co/MCM-41 were 343.78 m2/g, 0.026 cm3/g, and 16.442 Å, respectively. A decrease in the values of these parameters points toward the incorporation of nickel and cobalt, which results in the obstruction of the pore channel from the deposition of these species into the pores. A similar pattern was also observed by Gucbilmez et al. [38], who noted a reduction in the specific surface area and pore volume in V/MCM-41, with a rise in the ratio of vanadium to silicon. In another study by Ma et al. [39], they also observed a decrease in the surface area of Cu/MCM-41, with an increase in the percentage loading of copper over MCM-41.

3.2. Surface Morphology and Chemical Composition of Waste Sesame Biomass

SEM-EDX of sesame biomass in raw, demineralized, and catalyst incorporation was performed. The SEM pictures are presented in Figure 2a–c. Figure 2b contains tiny pores, which have been produced as a result of the treatment with the acid solution due to the removal of the minerals. Figure 2c indicates that the catalyst has effectively been distributed onto the surface of the demineralized sample [40].
Figure 2d–f shows the EDX of the raw, acid-treated, and catalyst-impregnated samples. The untreated sample shown in Figure 2a contains Zn, Cu, Fe, Ca, K, S, Si, P, Al, C, Mg, and Na. These minerals, especially the alkali metals, greatly affect the pyrolysis of the biomass, which act as a catalyst [41,42]; thus, it is essential to remove these minerals and study the influence of 9% Ni/8% Co/MCM-41 alone on the pyrolysis of sesame waste biomass. Figure 2e shows the EDX of the demineralized sample indicating complete removal of all the minerals after an acid wash. Figure 2f presents the loading of the acid-washed sample with 9% Ni/8% Co/MCM-41.

3.3. Thermogravimetric Analysis and Kinetic Study

In order to make a kinetic interpretation of the pyrolysis reaction, the experiments were performed with/without catalyst at various temperature programmed rates and the obtained results were used to calculate the kinetics parameters using the Kissinger method (Equation (1)). The subsequently constructed plots are presented in Figure 3a and b. The Ea and frequency factors were evaluated from the slope and intercept of the plots, and these are presented in Table 1. The results exhibit that pyrolysis of sesame biomass is a complex reaction, with many steps that mainly consist of the degradation of hemicellulose, cellulose, and lignin, while each step follows a typical behavior independent of each other [43]. These results are in harmony with previous reports. Ahmad, et al. [44] evaluated the pyrolysis kinetics of para grass using KAS and OFW equations and observed an Ea value from 103 to 233 kJ mol−1. Chen et al. [45] attributed the increase in Ea with %conversion to the decomposition of hemicellulose, cellulose, and lignin. Nisar et al. [46] carried out a thermogravimetric analysis of almond shells with/without zinc oxide and the decomposition was found to be a multistep reaction. Torres-García et al. [47] observed the decomposition of peanut shells as a multistep reaction involving the degradation of hemicellulose with an Ea of 172 kJ/mol, cellulose with an Ea of 203.4 kJ/mol, and lignin with an Ea of 218 kJ/mol using the Kissinger equation.

3.4. Pyrolysis of Biomass and Characterization of Bio-Oil

The influence of temperature on the product yields was checked by pyrolysis of biomass from 310 to 400 °C with catalyst for 40 min. The products formed are shown in Figure 4a. The figure shows the maximum bio-oil at 320 °C. An increase in gas and a decrease in biochar were noted with a rise in the temperature, which is because of the secondary degradation of the bio-oil [48,49]. Moreover, the effect of the time on the oil yield was also assessed at an optimized temperature, and an optimum time of 20 min was observed for the maximum oil yield, as shown in Figure 4b. The observations match well with previously reported studies [50]. Nizamuddin et al. [51] investigated the pyrolysis of biomass and an improvement in the percentage yield of the liquid fraction was noted with the elevation in temperature and the increase in time, up to a certain level, and then a reduction was noted. The trend has a similarity with our observation. GCMS was used for the analysis of bio-oil recovered from catalytic reactions at optimized conditions, as depicted by the chromatogram in Figure 5, whereas the detected compounds are provided in Table 2. The existence of phenols is ascribed to the degradation of lignin. Furfural production points towards the decomposition of hemicellulose, whereas the occurrence of different acids in the oil is because of the incomplete decomposition of hemicelluloses, cellulose, and lignin [52]. It can be illustrated from the table that the oil possesses mostly small molecules of furan, aldehydes, ketones, acids, and phenols [53]. It can also be observed from the GCMS analysis that the presence of valuable products, such as furfural, phenols, and maltol has increased the quality of the oil. Aromatics were also found in abundance, which is in conformity with the results of Zhang et al. [54]. Ateş et al. [55] studied the pyrolysis of sesame biomass from 400 to 700 °C. A maximum yield of 37.20 wt.% of bio-oil was obtained at 550 °C with alkenes, alkanes, and branched hydrocarbons in abundance. Jeon et al. [56] performed the catalytic decomposition of biomass and recovered a good quality of oil, compared to the uncatalyzed reaction.

3.5. Physicochemical Characteristics of Bio-Oil

The bio-oil from the catalytic decomposition of the waste sesame biomass had a dark brown color and pungent order. Its physicochemical properties, such as the viscosity, fluidity, density, and specific gravity and API gravity were determined and compared to other bio-oils to determine its characteristics (as shown in Table 3). From these properties, it was observed that the bio-oil is dependent on the nature and type of catalyst used. These observations agree with earlier reports. He et al. [57] investigated the physicochemical characteristics of bio-oil produced from switch grass and observed variations in the viscosity and density of the bio-oil with a change in the experimental conditions. The authors noted that there are no ideal conditions for bio-oil production with all the wanted characteristics and these must be adjusted according to the specific use of the bio-oil. Furthermore, a comparison of the different properties of the oil obtained in this work with previously reported studies (Table 3) indicates that the values of the viscosity and density of the bio-oil produced from the degradation of the sesame waste biomass in the presence of Ni/Co/MCM-41 are reduced compared to the oils obtained from the various sources of biomass in the presence/absence of different catalysts. Bio-oil with a higher viscosity and density results in inefficient atomization during the spraying of the oil, which leads to the accuracy of the fuel injectors becoming limited. Therefore, the oil produced from the decomposition of the sesame waste biomass in the presence of Ni/Co/MCM-41 has bright prospects in the replacement of commercial fuels.

4. Conclusions

MCM-41 was successfully prepared by employing a templating mechanism followed by 9% Ni/8% Co/MCM-41 using the sol-gel process. Characterization of the catalyst using SEM-EDX, XRD, and the surface area analyzer confirmed the successful incorporation of cobalt and nickel onto MCM-41, in the desired percentage. The prepared catalyst was used for the decomposition of sesame waste biomass in a pyrolysis chamber. The products obtained were studied through GC-MS. Thermogravimetry was used to determine the kinetics parameters. It has been concluded that 9% Ni/8% Co/MCM-41 reduced the maximum degradation temperature from 340 to 320 °C along with a reduction in the average activation energy, using the Kissinger model. The physicochemical characteristics of the oil were evaluated, and it was concluded that the bio-oil if upgraded properly, can be made equivalent to commercial fuel.

Author Contributions

Conceptualization, J.N.; Data curation, G.A.; Formal analysis, R.U. and G.A.; Investigation, R.U.; Methodology, R.U.; Project administration, J.N.; Resources, J.N.; Supervision, J.N.; Visualization, A.S., M.I.D. and Z.H.; Writing—original draft, R.U.; Writing—review & editing, J.N., G.A., A.S., M.I.D., Z.H. and R.A. All authors have read and agreed to the published version of the manuscript.

Funding

The support provided by Higher Education Commission, Pakistan through grant No. 20-1491.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yan, Y.; Wu, T.; Lester, E.; Tang, L.; Meng, Y.; Fang, Y.; Pang, C.H. The kinetics studies and thermal characterisation of biomass. Energy Procedia 2019, 158, 357–363. [Google Scholar] [CrossRef]
  2. Mishra, R.K.; Mohanty, K. Pyrolysis characteristics, fuel properties, and compositional study of Madhuca longifolia seeds over metal oxide catalysts. Biomass Convers. Biorefinery 2019, 10, 621–637. [Google Scholar] [CrossRef]
  3. Gupta, S.; Gupta, G.K.; Mondal, M.K. Slow pyrolysis of chemically treated walnut shell for valuable products: Effect of process parameters and in-depth product analysis. Energy 2019, 181, 665–676. [Google Scholar] [CrossRef]
  4. Dhyani, V.; Bhaskar, T. A comprehensive review on the pyrolysis of lignocellulosic biomass. Renew. Energy 2018, 129, 695–716. [Google Scholar] [CrossRef]
  5. Nisar, J.; Ullah, N.; Awan, I.A.; Iqbal, M.; Khan, T.A. Pyrolysis–gas chromatography of sugar beet bagasse. Waste Biomass Valorization 2016, 7, 79–85. [Google Scholar] [CrossRef]
  6. Nisar, J.; Nasir, U.; Ali, G.; Shah, A.; Farooqi, Z.H.; Iqbal, M.; Shah, M.R. Kinetics of pyrolysis of sugarcane bagasse: Effect of catalyst on activation energy and yield of pyrolysis products. Cellulose 2021, 28, 7593–7607. [Google Scholar] [CrossRef]
  7. Nisar, J.; Waris, S.; Shah, A.; Anwar, F.; Ali, G.; Ahmad, A.; Muhammad, F. Production of Bio-Oil from De-Oiled Karanja (Pongamia pinnata L.) Seed Press Cake Via Pyrolysis: Kinetics and Evaluation of Anthill as the Catalyst. Sustain. Chem. 2022, 3, 345–357. [Google Scholar] [CrossRef]
  8. Mishra, R.K.; Mohanty, K. Thermocatalytic conversion of non-edible Neem seeds towards clean fuel and chemicals. J. Anal. Appl. Pyrolysis 2018, 134, 83–92. [Google Scholar] [CrossRef]
  9. Salema, A.A.; Ting, R.M.W.; Shang, Y.K. Pyrolysis of blend (oil palm biomass and sawdust) biomass using TG-MS. Bioresour. Technol. 2019, 274, 439–446. [Google Scholar] [CrossRef] [PubMed]
  10. Shao, S.; Zhang, H.; Xiao, R.; Li, X.; Cai, Y. Controlled regeneration of ZSM-5 catalysts in the combined oxygen and steam atmosphere used for catalytic pyrolysis of biomass-derivates. Energy Convers. Manag. 2018, 155, 175–181. [Google Scholar] [CrossRef]
  11. Dabros, T.M.; Stummann, M.Z.; Høj, M.; Jensen, P.A.; Grunwaldt, J.-D.; Gabrielsen, J.; Mortensen, P.M.; Jensen, A.D. Transportation fuels from biomass fast pyrolysis, catalytic hydrodeoxygenation, and catalytic fast hydropyrolysis. Prog. Energy Combust. Sci. 2018, 68, 268–309. [Google Scholar] [CrossRef]
  12. Rehman, N.U.; Nisar, J.; Ali, G.; Ahmad, A.; Shah, A.; Farooqi, Z.H.; Muhammad, F. Production of Bio-Oil from Thermo-Catalytic Decomposition of Pomegranate Peels over a Sulfonated Tea Waste Heterogeneous Catalyst: A Kinetic Investigation. Energies 2023, 16, 1908. [Google Scholar] [CrossRef]
  13. Mohan, D.; Pittman, C.U., Jr.; Steele, P.H. Pyrolysis of wood/biomass for bio-oil: A critical review. Energy Fuels 2006, 20, 848–889. [Google Scholar] [CrossRef]
  14. Steele, P.; Puettmann, M.E.; Penmetsa, V.K.; Cooper, J.E. Life-cycle assessment of pyrolysis bio-oil production. For. Prod. J. 2012, 62, 326–334. [Google Scholar] [CrossRef]
  15. Nisar, J.; Sharaf, M.; Ali, G.; Farooqi, Z.; Iqbal, M.; Khan, S. Pyrolysis of juice-squeezed grapefruit waste: Effect of nickel oxide on kinetics and bio-oil yield. Int. J. Environ. Sci. Technol. 2022, 19, 10211–10222. [Google Scholar] [CrossRef]
  16. Rahman, M.M.; Liu, R.; Cai, J. Catalytic fast pyrolysis of biomass over zeolites for high quality bio-oil–a review. Fuel Process. Technol. 2018, 180, 32–46. [Google Scholar] [CrossRef]
  17. Ali, G.; Afraz, M.; Muhammad, F.; Nisar, J.; Shah, A.; Munir, S.; Hussain, S.T. Production of Fuel Range Hydrocarbons from Pyrolysis of Lignin over Zeolite Y, Hydrogen. Energies 2022, 16, 215. [Google Scholar] [CrossRef]
  18. Nisar, J.; Ahmad, A.; Ali, G.; Rehman, N.U.; Shah, A.; Shah, I. Enhanced Bio-Oil Yield from Thermal Decomposition of Peanut Shells Using Termite Hill as the Catalyst. Energies 2022, 15, 1891. [Google Scholar] [CrossRef]
  19. Adam, J.; Blazso, M.; Meszaros, E.; Stöcker, M.; Nilsen, M.H.; Bouzga, A.; Hustad, J.E.; Grønli, M.; Øye, G. Pyrolysis of biomass in the presence of Al-MCM-41 type catalysts. Fuel 2005, 84, 1494–1502. [Google Scholar] [CrossRef]
  20. Nilsen, M.H.; Antonakou, E.; Bouzga, A.; Lappas, A.; Mathisen, K.; Stöcker, M. Investigation of the effect of metal sites in Me–Al-MCM-41 (Me = Fe, Cu or Zn) on the catalytic behavior during the pyrolysis of wooden based biomass. Microporous Mesoporous Mater. 2007, 105, 189–203. [Google Scholar] [CrossRef]
  21. Antonakou, E.; Lappas, A.; Nilsen, M.H.; Bouzga, A.; Stöcker, M. Evaluation of various types of Al-MCM-41 materials as catalysts in biomass pyrolysis for the production of bio-fuels and chemicals. Fuel 2006, 85, 2202–2212. [Google Scholar] [CrossRef]
  22. Maria do Socorro, B.F.; Melo, D.M.; Fontes, L.A.; Braga, R.M.; Costa, C.C.; Martinelli, A.E. Ex situ catalytic biomass pyrolysis using mesoporous Ti-MCM-41. Environ. Sci. Pollut. Res. 2019, 26, 5983–5989. [Google Scholar]
  23. Li, X.; Hu, C.; Shao, S.; Cai, Y.; Zhang, X. In situ catalytic upgrading of pyrolysis vapours from vacuum pyrolysis of rape straw over La/MCM-41. J. Anal. Appl. Pyrolysis 2019, 140, 213–218. [Google Scholar] [CrossRef]
  24. Yu, L.; Farinmade, A.; Ajumobi, O.; Su, Y.; John, V.T.; Valla, J.A. MCM-41/ZSM-5 composite particles for the catalytic fast pyrolysis of biomass. Appl. Catal. A Gen. 2020, 602, 117727. [Google Scholar] [CrossRef]
  25. Xue, Z.; Zhong, Z.; Zhang, B. Microwave-Assisted Catalytic Fast Pyrolysis of Biomass for Hydrocarbon Production with Physically Mixed MCM-41 and ZSM-5. Catalysts 2020, 10, 685. [Google Scholar] [CrossRef]
  26. Ratnasari, D.K.; Yang, W.; Jönsson, P.G. Kinetic Study of an H-ZSM-5/Al–MCM-41 Catalyst Mixture and Its Application in Lignocellulose Biomass Pyrolysis. Energy Fuels 2019, 33, 5360–5367. [Google Scholar] [CrossRef]
  27. Li, K.; Wang, B.; Bolatibieke, D.-N.; Nan, D.-H.; Zhang, Z.-X.; Cui, M.-S.; Lu, Q. Catalytic fast pyrolysis of biomass with Ni-P-MCM-41 to selectively produce levoglucosenone. J. Anal. Appl. Pyrolysis 2020, 148, 104824. [Google Scholar] [CrossRef]
  28. Blaine, R.L.; Kissinger, H.E. Homer Kissinger and the Kissinger equation. Thermochim. Acta 2012, 540, 1–6. [Google Scholar] [CrossRef]
  29. Grün, M.; Unger, K.K.; Matsumoto, A.; Tsutsumi, K. Novel pathways for the preparation of mesoporous MCM-41 materials: Control of porosity and morphology. Microporous Mesoporous Mater. 1999, 27, 207–216. [Google Scholar] [CrossRef]
  30. Nejat, T.; Jalalinezhad, P.; Hormozi, F.; Bahrami, Z. Hydrogen production from steam reforming of ethanol over Ni-Co bimetallic catalysts and MCM-41 as support. J. Taiwan Inst. Chem. Eng. 2019, 97, 216–226. [Google Scholar] [CrossRef]
  31. Wu, C.; Wang, L.; Williams, P.T.; Shi, J.; Huang, J. Hydrogen production from biomass gasification with Ni/MCM-41 catalysts: Influence of Ni content. Appl. Catal. B Environ. 2011, 108, 6–13. [Google Scholar] [CrossRef]
  32. Parangi, T.F.; Patel, R.M.; Chudasama, U.V. Synthesis and characterization of mesoporous Si-MCM-41 materials and their application as solid acid catalysts in some esterification reactions. Bull. Mater. Sci. 2014, 37, 609–615. [Google Scholar] [CrossRef]
  33. Abrokwah, R.Y.; Deshmane, V.G.; Kuila, D. Comparative performance of M-MCM-41 (M: Cu, Co, Ni, Pd, Zn and Sn) catalysts for steam reforming of methanol. J. Mol. Catal. A Chem. 2016, 425, 10–20. [Google Scholar] [CrossRef]
  34. Salam, M.S.A.; Betiha, M.A.; Shaban, S.A.; Elsabagh, A.M.; Abd El-Aal, R.M. Synthesis and characterization of MCM-41-supported nano zirconia catalysts. Egypt. J. Pet. 2015, 24, 49–57. [Google Scholar] [CrossRef]
  35. Singh, S.B.; De, M. Scope of doped mesoporous (<10 nm) surfactant-modified alumina templated carbons for hydrogen storage applications. Int. J. Energy Res. 2019, 43, 4264–4280. [Google Scholar]
  36. Parvulescu, V.; Su, B.-L. Iron, cobalt or nickel substituted MCM-41 molecular sieves for oxidation of hydrocarbons. Catal. Today 2001, 69, 315–322. [Google Scholar] [CrossRef]
  37. Liu, D.; Quek, X.Y.; Cheo, W.N.E.; Lau, R.; Borgna, A.; Yang, Y. MCM-41 supported nickel-based bimetallic catalysts with superior stability during carbon dioxide reforming of methane: Effect of strong metal–support interaction. J. Catal. 2009, 266, 380–390. [Google Scholar] [CrossRef]
  38. Gucbilmez, Y.; Dogu, T.; Balci, S. Vanadium incorporated high surface area MCM-41 catalysts. Catal. Today 2005, 100, 473–477. [Google Scholar] [CrossRef]
  39. Ma, X.; Chi, H.; Yue, H.; Zhao, Y.; Xu, Y.; Lv, J.; Wang, S.; Gong, J. Hydrogenation of dimethyl oxalate to ethylene glycol over mesoporous Cu-MCM-41 catalysts. AIChE J. 2013, 59, 2530–2539. [Google Scholar] [CrossRef]
  40. Nisar, J.; Ali, F.; Malana, M.A.; Ali, G.; Iqbal, M.; Shah, A.; Bhatti, I.A.; Khan, T.A.; Rashid, U. Kinetics of the pyrolysis of cobalt-impregnated sesame stalk biomass. Biomass Convers. Biorefinery 2020, 10, 1179–1187. [Google Scholar] [CrossRef]
  41. Balonek, C.M.; Lillebø, A.H.; Rane, S.; Rytter, E.; Schmidt, L.D.; Holmen, A. Effect of alkali metal impurities on Co–Re catalysts for Fischer–Tropsch synthesis from biomass-derived syngas. Catal. Lett. 2010, 138, 8–13. [Google Scholar] [CrossRef]
  42. Zhu, W.; Song, W.; Lin, W. Catalytic gasification of char from co-pyrolysis of coal and biomass. Fuel Process. Technol. 2008, 89, 890–896. [Google Scholar] [CrossRef]
  43. Aburto, J.; Moran, M.; Galano, A.; Torres-García, E. Non-isothermal pyrolysis of pectin: A thermochemical and kinetic approach. J. Anal. Appl. Pyrolysis 2015, 112, 94–104. [Google Scholar] [CrossRef]
  44. Ahmad, M.S.; Mehmood, M.A.; Al Ayed, O.S.; Ye, G.; Luo, H.; Ibrahim, M.; Rashid, U.; Nehdi, I.A.; Qadir, G. Kinetic analyses and pyrolytic behavior of Para grass (Urochloa mutica) for its bioenergy potential. Bioresour. Technol. 2017, 224, 708–713. [Google Scholar] [CrossRef]
  45. Chen, Z.; Hu, M.; Zhu, X.; Guo, D.; Liu, S.; Hu, Z.; Xiao, B.; Wang, J.; Laghari, M. Characteristics and kinetic study on pyrolysis of five lignocellulosic biomass via thermogravimetric analysis. Bioresour. Technol. 2015, 192, 441–450. [Google Scholar] [CrossRef]
  46. Nisar, J.; Rahman, A.; Ali, G.; Shah, A.; Farooqi, Z.H.; Bhatti, I.A.; Iqbal, M.; Rehman, N.U. Pyrolysis of almond shells waste: Effect of zinc oxide on kinetics and product distribution. Biomass Convers. Biorefinery 2020, 12, 2583–2595. [Google Scholar] [CrossRef]
  47. Torres-García, E.; Ramírez-Verduzco, L.; Aburto, J. Pyrolytic degradation of peanut shell: Activation energy dependence on the conversion. Waste Manag. 2020, 106, 203–212. [Google Scholar] [CrossRef]
  48. Westerhof, R.J.; Brilman, D.W.; van Swaaij, W.P.; Kersten, S.R. Effect of temperature in fluidized bed fast pyrolysis of biomass: Oil quality assessment in test units. Ind. Eng. Chem. Res. 2010, 49, 1160–1168. [Google Scholar] [CrossRef]
  49. Demirbas, A. Effect of temperature on pyrolysis products from four nut shells. J. Anal. Appl. Pyrolysis 2006, 76, 285–289. [Google Scholar] [CrossRef]
  50. Zhang, S.; Yan, Y.; Li, T.; Ren, Z. Upgrading of liquid fuel from the pyrolysis of biomass. Bioresour. Technol. 2005, 96, 545–550. [Google Scholar] [CrossRef]
  51. Nizamuddin, S.; Baloch, H.A.; Mubarak, N.M.; Riaz, S.; Siddiqui, M.T.H.; Takkalkar, P.; Tunio, M.; Mazari, S.; Bhutto, A.W. Solvothermal liquefaction of corn stalk: Physico-chemical properties of bio-oil and biochar. Waste Biomass Valorization 2019, 10, 1957–1968. [Google Scholar] [CrossRef]
  52. Uçar, S.; Karagöz, S. The slow pyrolysis of pomegranate seeds: The effect of temperature on the product yields and bio-oil properties. J. Anal. Appl. Pyrolysis 2009, 84, 151–156. [Google Scholar] [CrossRef]
  53. Chen, D.; Zhou, J.; Zhang, Q. Effects of torrefaction on the pyrolysis behavior and bio-oil properties of rice husk by using TG-FTIR and Py-GC/MS. Energy Fuels 2014, 28, 5857–5863. [Google Scholar] [CrossRef]
  54. Zhang, B.; Zhong, Z.; Min, M.; Ding, K.; Xie, Q.; Ruan, R. Catalytic fast co-pyrolysis of biomass and food waste to produce aromatics: Analytical Py–GC/MS study. Bioresour. Technol. 2015, 189, 30–35. [Google Scholar] [CrossRef]
  55. Ateş, F.; Pütün, E.; Pütün, A.E. Fast pyrolysis of sesame stalk: Yields and structural analysis of bio-oil. J. Anal. Appl. Pyrolysis 2004, 71, 779–790. [Google Scholar] [CrossRef]
  56. Jeon, M.-J.; Jeon, J.-K.; Suh, D.J.; Park, S.H.; Sa, Y.J.; Joo, S.H.; Park, Y.-K. Catalytic pyrolysis of biomass components over mesoporous catalysts using Py-GC/MS. Catal. Today 2013, 204, 170–178. [Google Scholar] [CrossRef]
  57. He, R.; Ye, X.P.; English, B.C.; Satrio, J.A. Influence of pyrolysis condition on switchgrass bio-oil yield and physicochemical properties. Bioresour. Technol. 2009, 100, 5305–5311. [Google Scholar] [CrossRef]
  58. Guo, X.; Wang, S.; Wang, Q.; Guo, Z.; Luo, Z. Properties of bio-oil from fast pyrolysis of rice husk. Chin. J. Chem. Eng. 2011, 19, 116–121. [Google Scholar] [CrossRef]
  59. Bakar, M.S.A.; Titiloye, J.O. Catalytic pyrolysis of rice husk for bio-oil production. J. Anal. Appl. Pyrolysis 2013, 103, 362–368. [Google Scholar] [CrossRef]
  60. Lazzari, E.; Schena, T.; Primaz, C.T.; da Silva Maciel, G.P.; Machado, M.E.; Cardoso, C.A.L.; Jacques, R.A.; Caramão, E.B. Production and chromatographic characterization of bio-oil from the pyrolysis of mango seed waste. Ind. Crops Prod. 2016, 83, 529–536. [Google Scholar] [CrossRef]
  61. Grioui, N.; Halouani, K.; Agblevor, F.A. Bio-oil from pyrolysis of Tunisian almond shell: Comparative study and investigation of aging effect during long storage. Energy Sustain. Dev. 2014, 21, 100–112. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) MCM-41 and (b) Ni-Co/MCM-41; EDX of (c) MCM-41 and (d) Ni-Co/MCM-41; XRD of (e) MCM-41 and (f) Ni-Co/MCM-41.
Figure 1. SEM images of (a) MCM-41 and (b) Ni-Co/MCM-41; EDX of (c) MCM-41 and (d) Ni-Co/MCM-41; XRD of (e) MCM-41 and (f) Ni-Co/MCM-41.
Energies 16 03731 g001aEnergies 16 03731 g001b
Figure 2. SEM of (a) raw, (b) demineralized, and (c) catalyst-incorporated and EDX of (d) raw, (e) after acid wash, and (f) sesame waste biomass impregnated with catalyst.
Figure 2. SEM of (a) raw, (b) demineralized, and (c) catalyst-incorporated and EDX of (d) raw, (e) after acid wash, and (f) sesame waste biomass impregnated with catalyst.
Energies 16 03731 g002aEnergies 16 03731 g002bEnergies 16 03731 g002c
Figure 3. Kissinger plots for (a) SWB without catalyst and (b) with catalyst.
Figure 3. Kissinger plots for (a) SWB without catalyst and (b) with catalyst.
Energies 16 03731 g003
Figure 4. Temperature optimization. (a) Temperature effect and (b) time optimization at optimized temperature.
Figure 4. Temperature optimization. (a) Temperature effect and (b) time optimization at optimized temperature.
Energies 16 03731 g004aEnergies 16 03731 g004b
Figure 5. GCMS of bio–oil from catalytic decomposition of sesame waste biomass.
Figure 5. GCMS of bio–oil from catalytic decomposition of sesame waste biomass.
Energies 16 03731 g005
Table 1. Kinetic parameters were calculated using the Kissinger model with/without a catalyst.
Table 1. Kinetic parameters were calculated using the Kissinger model with/without a catalyst.
ComponentsCatalyticNon-Catalytic
Ea (kJ/mol)A (min−1)Ea (kJ/mol)A (min−1)
Hemicellulose133.022.3 × 10891.458.0 × 105
Cellulose141.338.0 × 10999.763.8 × 106
Lignin191.221.1 × 1011149.651.6 × 109
Table 2. Compounds traced in GCMS of bio-oil recovered from the catalyzed decomposition of sesame waste.
Table 2. Compounds traced in GCMS of bio-oil recovered from the catalyzed decomposition of sesame waste.
Peak No.R/Time
(min)
Compound NameMol.
Formula
Mol. wt.Percent Area
11.85FurfuralC5H4O29619.35
22.73Cyclopropane,2-(1,1-dimethyl-2-pentenyl)-1,1-dimethylC12H221665.78
33.302-Furancarboxaldehyde, 5-methylC6H6O21109.94
44.221,2-Cyclopentanedione, 3-methylC6H8O211210.70
55.01Phenol, 2-methoxy-C7H8O21247.91
65.40MaltolC6H6O31263.72
75.872-Myristynoic acidC14H24O22240.48
86.44Phenol, 2-methoxy-4-methyl-C8H10O21382.89
96.781,4:3,6-Dianhydro-à-d-glucopyranoseC6H8O41443.06
107.132-Furancarboxaldehyde,
5-(hydroxymethyl)
C6H6O31264.92
117.60Phenol, 4-ethyl-2-methoxy-C9H12O21522.28
128.052-Oxabicyclo[3.3.0]oct-7-en-3-one,
7-(1-hydroxypentyl)-
C12H18O32101.19
148.60Phenol, 2,6-dimethoxy-C8H10O31546.94
149.21VanillinC8H8O31523.22
159.781,2,4-TrimethoxybenzeneC9H12O31682.28
1610.29Ethanone,
1-(4-hydroxy-3-methoxyphenyl)-
C9H10O31661.29
1710.80Silane, dimethoxymethylphenylC9H14O2Si1823.37
1811.63Phenol,2,6-dimethoxy-4-(2propenyl)-C11H14O31941.69
1912.28Benzaldehyde,4-hydroxy-3,5-dimethoxy-C9H10O41823.75
2013.08Ethanone,1-(4-hydroxy-3,5-dimethoxyphenyl)-C10H12O41962.33
2113.473,5-Dimethoxy-4 hydroxyphenylaceticAcidC10H12O52121.34
2214.02Ethanone,1-(4-hydroxy-3,5-dimethoxyphenyl)-C10H12O41960.58
2314.553,9;7,8-Diepoxybicyclo[4.3.0]nonane-6á-ol-2-one,9à-acetoxy-4-isopropenyl-5-methoxycarbonyl-1-methyl-C18H22O83660.15
2414.897,9-Di-tert-butyl-1-oxaspiro(4,5)deca-6,
9-diene-2,8-dione
C17H24O32760.32
2515.28Estra-1,3,5(10)-trien-17á-olC18H24O2560.32
2615.563,5-Dimethoxy-4-hydroxycinnamaldehyC11H12O42080.88
2716.09Curan-19,20-diol, 16,17-didehydro-,
(19S)-
C19H24N2O23120.07
2816.528-Octadecenoic acid, methyl esterC19H36O22960.07
2917.17Octadecanoic acidC18H36O22840.38
3018.582-Nonadecanone
2,4-dinitrophenylhydrazine
C25H42N4O44620.02
3118.939,12,15-Octadecatrienoic acid,
2,3-bis[(trimethylsilyl)oxy]propyl ester,(Z,Z,Z)-
C27H52O4Si24960.05
3220.211,2-Benzenedicarboxylicacid,diisooctyl esterC24H38O43900.32
3321.782,5-Dimethoxy-4-ethylamphetamineC13H21NO22230.12
3426.35SertindoleC24H26ClFN4O4400.43
Table 3. Comparison of physicochemical characteristics of bio-oil obtained from different sources.
Table 3. Comparison of physicochemical characteristics of bio-oil obtained from different sources.
BiomassViscosity (cP)Density
gm−3
Fluidity
(cP)−1
Specific GravityAPI GravityReference
Rice husk82.431.190---[58]
Rice husk/Al-MCM-411.65----[59]
Mango seed149----[60]
Almond shell<100----[61]
Cobalt-impregnated sesame stalk1761.047 [40]
Sesame waste biomass/Ni/Co/MCM-411.131.010.9751.231−1.687This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nisar, J.; Ullah, R.; Ali, G.; Shah, A.; Din, M.I.; Hussain, Z.; Amin, R. Thermocatalytic Decomposition of Sesame Waste Biomass over Ni-Co-Doped MCM-41: Kinetics and Physicochemical Properties of the Bio-Oil. Energies 2023, 16, 3731. https://doi.org/10.3390/en16093731

AMA Style

Nisar J, Ullah R, Ali G, Shah A, Din MI, Hussain Z, Amin R. Thermocatalytic Decomposition of Sesame Waste Biomass over Ni-Co-Doped MCM-41: Kinetics and Physicochemical Properties of the Bio-Oil. Energies. 2023; 16(9):3731. https://doi.org/10.3390/en16093731

Chicago/Turabian Style

Nisar, Jan, Raqeeb Ullah, Ghulam Ali, Afzal Shah, Muhammad Imran Din, Zaib Hussain, and Roohul Amin. 2023. "Thermocatalytic Decomposition of Sesame Waste Biomass over Ni-Co-Doped MCM-41: Kinetics and Physicochemical Properties of the Bio-Oil" Energies 16, no. 9: 3731. https://doi.org/10.3390/en16093731

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop