banner
Home / News / Efficient biodiesel production from oleic and palmitic acid using a novel molybdenum metal
News

Efficient biodiesel production from oleic and palmitic acid using a novel molybdenum metal

Jan 26, 2024Jan 26, 2024

Scientific Reports volume 12, Article number: 10338 (2022) Cite this article

2873 Accesses

5 Citations

17 Altmetric

Metrics details

In this study, metal–organic framework based on molybdenum and piperidine-4-carboxylic acid, was synthesized through a simple solvothermal method and employed as an effective catalyst for biodiesel production from oleic acid and palmitic acid via esterification reaction. The prepared catalyst was characterized by XRD, FTIR, TGA, DSC, BET, SEM, TEM, ICP-OES, X-ray mapping and EDX analysis. The resulting Mo–MOF catalyst exhibit a rod-like morphology, specific surface area of 56 m2/g, and thermal stability up to 300 °C. The solid catalyst exhibited high activities for esterification of oleic acid and palmitic acid. Moreover, the catalyst could be simply recovered and efficiently reutilized for several times without significant loss in its activity, also obtained results revealed that metal–organic framework could be used for the appropriate and rapid biodiesel production.

Due to increasing environmental pollution, global warming caused by fossil fuels, many studies are seeking for develop renewable energy technologies. In this context, biodiesel are considered to have the most potential to reduce the amount of particulate matter1, CO22, and greenhouse gas emissions3 due to high octane number4, and low viscosity5 that makes it a promising and economically feasible alternative to common fossil fuels6. Biodiesel is a renewable energy that can be produced by transesterification of triglyceride in biomass-oil (such as vegetable oil and animal fat) with methanol7. Traditional biodiesel production methods utilize homogeneous conditions in the presence of bases or an acid catalyst such as sulfuric acid and sodium hydroxide8.These homogeneous systems suffer from limitations such as corrosion of reactors, difficult recovery and recyclability of catalyst and environmental pollution9,10. In this context, heterogeneous catalysts provide more advantages over homogeneous catalysts, such as recyclable10,11, ease separation-purification process12, high glycerol purity13, and not corrosive14. To date, numerous solid acid and base catalysts with tunable structural and surface functionalities have been offered, and many of them shows high catalytic activity for biodiesel production yield. Heterogeneous solid base catalysts usually provide higher rates compared with the acid counterparts under milder reaction conditions. However, they cannot be directly employed for oil having more than 2 wt% FFA due to side reactions, such as saponification and hydrolysis, and decrease both the catalyst activity and the ester yield15. Hence, solid acid catalysts are applied when dealing with low-quality or non-edible vegetable oils containing significant quantities of FFAs and water15. Molybdenum compounds have been recognized as versatile catalysts, because of the ability of this metal to be on the solid surface in different oxidation states, ranging from Mo6+ to metallic Mo (Mo0)16.

Anhydrous sodium molybdate17, bulk MoO318, silica MoO3/B-ZSM-519 molybdenum supported on alumina20, silica, silica-alumina, and titania21,22, as well as carbon23 have been used as esterification and transesterification catalysts for biodiesel production from several oils, including waste oil. Gandía et al., described the application of bulk and Al2O3-supported molybdenum oxide for the production of biodiesel from Oil. Control experiments showed that bulk MoO3 is very active for both transesterification and esterification reactions, but it suffered from severe molybdenum leaching in the reaction medium.

Compared to bulk MoO3, alumina-supported MoO3 leads to a more efficient utilization of the active phase and enhanced stability towards molybdenum leaching by the reaction medium24. In this study, we have been introduced a novel MOF as a highly efficient separable catalyst for biodiesel production from oleic acid and palmitic acid via an esterification reaction. In recent years, metal–organic frameworks (MOFs) are attracting increasing attention due to their important properties such as controllable composition25, large surface area26,27, thermal stability28, flexibility, and easy preparation29. MOFs are constructed from SBU connected by organic linkers to form extended coordination networks. The linkers widely used in MOFs are rigid organic chelators such as amino acid, terephthalic acid, and polycarboxylate ligands. Many factors affect the activity of MOFs, such as the organic ligand, solvent type, particle size, and metal sort. Metal–organic frameworks (MOFs) known as coordination polymers, undergoing study by many researchers for advanced applications, including such as catalysis30,31,32, separation33,34, gas storage35, carbon dioxide capture36. MOFs mainly have their adjustable nano-structures and porous properties. However, as a good carrier, MOF also has its inherent catalytic performance. Moreover, the MOF-based functional catalytic materials show great potential in biodiesel production and other relevant biorefineries. In Table 1, typical heterogeneous bifunctional catalysts were chosen to compare with MOF-based bifunctional catalysts for biodiesel production. These studies revealed that the MOF with amino groups (Brønsted base) showed high catalytic activity and mild conditions for biodiesel production than another heterogeneous acid–base catalyst (Table 1).

To the best of our knowledge, there have been few papers referred to preparation of Mo–MOFs47,48,49,50,51,52. In this study, the catalytic performance of metal–organic framework based on molybdenum and piperidine-4-carboxylic acid was employed in biodiesel production through the esterification of oleic acid and palmitic acid with methanol (Fig. 1).

The application of metal–organic framework for biodiesel production.

The structure of a network is affected by several factors such as building blocks, solvent, temperature, pH, and so on. Also, its topology is mainly dependent on the connectivity and the symmetry of the metal ions (or metal clusters) and organic nodes. We find the rod-MOF topology [Rod MOFs are metal–organic frameworks in which the metal-containing secondary building units consist of infinite rods of linked metal-centered polyhedral], according to the present approach, the metal ions are connected to the framework by the organic SBUs (Scheme 1).

The topological structure of Mo–MOF.

Mo–MOF was prepared by mixing Na2MoO4 with 4-piperidinecarboxylic acid (2:1 molar ratio M:L) in 20 mL of the DMF/H2O (18:2; v/v) and then magnetically stirred at 200 rpm for a 30 min at room temperature. Subsequently, the mixture was transferred to a Teflon-lined stainless steel autoclave and heated isothermally at 170 °C for 24 h. After, the autoclave was gradually cooled to room temperature. Finally, the resultant brown crystal was filtered by vacuum filtration followed by washing with ethyl acetate and dried at 60 °C in vacuum oven for 6 h (Scheme 1).

The catalytic activity of Mo–MOF was assessed for esterification reactions of oleic acid and palmitic acid by mixing oil (1 mol), methanol (13 mol), and MOF (300 mg) mixed in a round-bottom flask. The mixture was heated at 60 °C for 4 h. After the reaction is done, the catalyst was separated by centrifuges (5000 rmp), and the excess methanol was removed from the upper liquid phase using rotary evaporation completely. The extracted organic phase was further washed with distilled water to remove the residual impurities by decantation and dried with anhydrous Na2SO4.

The catalytic system has been studied by FTIR, XRD, BET, SEM, TEM, EDX, ICP-OES, TGA, DSC, and EDS elemental mapping. Moreover, Fourier transforms infrared (FT-IR) spectroscopy has been introduced for the quantitative analysis of Mo-based metal–organic frameworks. Figure 2 provides the FT-IR spectra of Na2MoO4. 2H2O (a), 4-piperidinecarboxylic acid (b), Mo-based metal–organic framework (c). FT-IR spectrum of 4-piperidine carboxylic acid exhibited two peaks at 3525 and 3450 cm−1 corresponding to stretching vibrations of –OH and –NH groups, respectively. The peak at 1645 cm−1 was attributed to the carboxylic (-COO) group of piperidine. The peaks at 2854 and 2967 cm−1 are corresponded to –C–H stretching of the methylene group. The strong peak of C–N stretching was observed at 1406 cm-1. Moreover, –OH bending and –N–H wagging of secondary amine were observed at 970 and 686 cm−1, respectively (Fig. 2a). The stretching vibration of Fig. 2b, was detected as a strong Mo–O stretch in the [MoO4]2− tetrahedrons at 824–634 cm−1 and additional weak peaks of the Mo–O bending mode around 500 cm−1. Moreover, the band at 3439 cm−1 is the characteristic of the stretching modes of the O–H bond of Na2MoO4. 2H2O. It can be observed that the carboxylic (−COO) group of piperidine carboxylic acid in the Mo-based metal–organic framework at 1606 cm−1 disappeared, indicating that the attachment of the carboxylic (−COO) group to the metal (Fig. 2c). It is worth noting that the C–N stretch band, which was present at 1406 cm−1 in the spectrum of 4-piperidine carboxylic acid, which is absent in the spectrum of the Mo-based metal–organic framework (Fig. 2c).

FT-IR spectra of 4-piperidinecarboxylic acid (a) Na2MoO4·2H2O (b) Mo-based metal organic framework, before recovery (c) Mo-based metal organic framework, after recovery (d).

Powder X-ray diffraction was used to determine the chemical composition and crystal structure of the typical synthesized products. Obviously, Mo–MOF exhibits a series of sharp peaks, indicating its good crystallization. The characteristic peaks are at 28.17, 33.12, 49.37, 52.7, 57.2, 65.17, and 68.2, which are identical to representative reference47,48, (Fig. 3). All of these characteristics indicate that Mo–MOF was successfully synthesized.

XRD pattern of Mo-based metal–organic framework (MOF).

Figure 4 shows the X-ray mapping and EDX analysis of Mo–MOF. The EDX spectrum indicates the percentage of index elements in Mo–MOF (C = 9.38%, N = 4.14%, O = 43.61% and Mo = 42.87%). Mapping analysis, reveals the occurrence of Mo as metallic and C, N, O as non-metallic constituents are homogeneously distributed within the metal–organic framework (MOF).

X-ray mapping analysis and EDX spectra for Mo–MOF.

The morphologies of Mo–MOF are monitored by scanning electron microscopy. The SEM images (Fig. 5) revealed that morphology of the product featured a rod-like structure with a well-defined plane facing along the transversal direction at approximately 500 nm–2 µm. The inset in Fig. 5a displays the edge of a bar with the typical cleavage for monoclinic structures (inset Fig. 5a).

SEM images of Mo-metal–organic framework (MOF).

The structure and morphology of the Mo-metal–organic framework (MOF) have been studied by TEM. Figure 6 shows that the particles have a nanorod-like structure.

TEM images of Mo-metal–organic framework (MOF).

TGA (Thermo-gravimetric analysis) was performed to investigate the thermal stability of the synthesized materials and was carried out from room temperature to 1000 K by using a TGA analyzer at a heating rate of 10 K/min under air atmosphere (Fig. 7). The result demonstrated thermodynamic stability of the materials up to 300 °C. The weight loss was observed in two steps: the first weight loss (3%) occurred in the range of 100 °C related to the loss of physically adsorbed H2O molecules and organic solvents. The second weight loss (19%) between 330 and 600 °C is corresponding to the decomposition of the organic species. The DSC shows two endotherms at 270 and 330 °C which corresponds to the successive release of water and DMF molecules from the host channels. The DSC analysis also displays an exothermic after complete loss of functional groups of guest molecules and coordination modes at approximately 420 °C which show that the compound undergoes phase transformation.

TGA of Mo-metal–organic framework (MOF).

The texture and porosity of the Mo–MOF was quantified by measuring the nitrogen adsorption isotherm (Fig. 8). The results show that according to the IUPAC classification of adsorption isotherms, the N2 isotherm resembles the type III having sharp adsorption capacity while indicating the presence of broader pore size distributions, and narrower mesopores, and wider micropores. According to The Brunauer–Emmett–Teller (BET) method, the specific surface area and pore volume were estimated to be 56 m2g−1 and 12. 98 cm3g−1. The BJH pore size calculations using the adsorption branch of the nitrogen isotherm indicate a micropore peak at about 1.66 nm of diameter for Mo–MOF (Fig. 9).

Nitrogen adsorption–desorption isotherms of Mo–MOF.

BJH pore size distribution for Mo–MOF.

In this study, the catalytic properties of the investigated Mo–MOF systems were performed by the esterification of oleic acid and palmitic acid with methanol. The biodiesel production was optimized using Mo–MOF as catalyst (100–300 mg), and various oleic acid/alcohol molar ratios at different temperatures. According to the obtained results, the yield of the reaction decreased with the catalyst amount decreasing. This may be due to the decrease in the total sorption surface area available to adsorbates resulting from the overlapping of active sites. According to these results, the maximum biodiesel production was achieved by300 mg of catalyst. A maximum conversion of 95% (oil to ester) was achieved for the temperature of 60 °C. The molar ratio between methanol and oil was considered to be 13:1 in this study for the completion of the esterification process (Table 2).

The physicochemical properties of the biodiesel, such as density, viscosity, flash point, pour points, and ash was determined according to the standard ASTM methods. The resulting biodiesel was pure and showed excellent physical and chemical properties meeting international standards. Kinematic viscosity is both cold flow and critical property of an oil to be used in compression engines and it represents the degree of resistance to flow offered by the fluid. The viscosity of biodiesel should be in the range of 1.9–6.0 mm2/s. The Kinematic viscosity of biodiesels made from palmitic acid and oleic acid are calculated to be 3.9 and 4.06 °C, respectively, which indicates these parameters have met the appropriate standard of ASTM D445. The flashpoint is the lowest temperature at which a liquid can form an ignitable mixture in air near the surface of the liquid. From Table 3, the flashpoint of biodiesels made from palmitic acid and oleic acid are measured to be 140, and 160 °C, respectively which is in agreement with ASTM D92. The pour point represents the lowest temperature at which a liquid will begin to flow. The pour point of biodiesels made from palmitic acid and oleic acid is calculated to be 12 and − 3 °C, respectively, which agrees with ASTM D97. The cloud point is the temperature at which wax crystals begin to form in a liquid as it is cooled. The cloud point of the biodiesel from palmitic acid and oleic acid were 17 and − 1 °C, respectively. Cloud points of biodiesels were almost within the ASTM D2500 standard range. Ash content: when organic compounds are decomposed at high temperatures (500–600 °C), the leftover residue is called ash. The approved limits of the carbon and ash residue after biofuel ignition are 0.01 wt% according to ASTM D-482, respectively. The obtained ash residues of the biodiesel from palmitic acid and oleic acid were between 0.004 and 0.002 wt% using Mo–MOF. These values are comparatively lower than those obtained in the case of petroleum fuel.

In order to show efficiency of the catalytic activity of Mo–MOF, we compared our results for the esterification of reaction of oleic acid has been shown in Table 4. It is evident that our protocol shows excellent catalytic activity in terms of yield of product.

Reusability emerged as a very important factor to determine the robustness of a heterogeneous solid catalyst and its commercial potential, particularly from economic and practical viewpoints. Filtration and centrifugation are two of the methods traditionally employed on a laboratory scale to allow handling, separation, recovery, and recycling of heterogeneous catalysts. Furthermore, the MOFs are often suitable as highly recyclable catalytic systems with easy and efficient recovery of catalysts through simple filtration or centrifugation. In this regard, the catalytic stability of Mo–MOF was investigated by the esterification of oleic acid. After the reaction was completed, the catalyst was separated using centrifugation. Then, it was washed with methanol and dried entirely at 80 °C. The dried catalyst was reused in the second reaction under optimum reaction conditions (Fig. 10). The catalyst was recycled efficiently for four cycles and the catalyst showed almost constant activity in the esterification of oleic acid (92% after the 4th recycle was obtained). The retrieved catalyst after four recycles maintained its original framework structure, as evidenced by FT-IR technique. The FT-IR spectrum of the recovered Mo–MOF indicates that this catalyst can be recycled without any change in its structure (Fig. 2d). Moreover, the nature of the catalyst was also determined by investigating the leaching of the molybdenum species by using ICP-AES analysis. ICP-OES analysis exhibited very low leaching of molybdenum species upon reuse. Based on the results from the ICP-OES analysis, the amount of molybdenum in the fresh catalyst and the recovered catalyst after four runs was 0.077 mol g−1 and 0.072 mol g−1 respectively, which this observation was attributed to the strong interactions of heteroatoms in the piperidine-4-carboxylic acid with molybdenum. Moreover, the results confirmed the heterogeneous nature of the catalysis.

Recyclability of Mo–MOF in the esterification of oleic acid with methanol.

In summary, a novel Mo–MOF was prepared by solvothermal method. Structural features of the material were ascertained by several analytical techniques like XRD, FTIR, TGA, DSC, BET, SEM, TEM, ICP-OES, X-ray mapping, and EDX analysis. Interestingly, obtained solid catalyst exhibits more highly catalytic activity in the esterification of oleic acid and palmitic acid with methanol. The influential reaction parameters, including the methanol/oil molar ratio, temperature, and catalyst amount, are optimized. The physicochemical properties of the biodiesel, such as density, viscosity, flash point, pour points, and ash was determined according to the standard ASTM methods. Control experiments showed that the fuel properties of prepared biodiesel are found to meet the international biodiesel standards. Moreover, the catalyst could be simply recovered and efficiently reutilized several times without significant loss in its activity, also obtained results revealed that this metal–organic framework could be used for the appropriate and rapid biodiesel production. This strategy can produce a solid catalyst with catalytic site accessibility, which has a good application future in the field of green chemistry.

Kegl, B. NOx and particulate matter (PM) emissions reduction potential by biodiesel usage. Energy. Fuels. 21, 3310–3316 (2007).

Article CAS Google Scholar

Thiyagarajan, S., Geo, V. E., Martin, L. J. & Nagalingam, B. Comparative analysis of various methods to reduce CO2 emission in a biodiesel fueled CI engine. Fuel 253, 146–158 (2019).

Article CAS Google Scholar

Esteves, E. M. M., Esteves, V. P. P., Bungenstab, D. J., Araújo, O. D. Q. F. & Morgado, C. D. R. V. Greenhouse gas emissions related to biodiesel from traditional soybean farming compared to integrated crop-livestock systems. J. Clean. Prod. 179, 81–92 (2018).

Article CAS Google Scholar

Han, W. Q. & Yao, C. D. Research on high cetane and high octane number fuels and the mechanism for their common oxidation and auto-ignition. Fuel 150, 29–40 (2015).

Article CAS Google Scholar

Knothe, G. & Steidley, K. R. Kinematic viscosity of biodiesel fuel components and related compounds: Influence of compound structure and comparison to petrodiesel fuel components. Fuel 84, 1059–1065 (2005).

Article CAS Google Scholar

Mosarof, M. H. et al. Implementation of palm biodiesel based on economic aspects, performance, emission, and wear characteristics. Energy Convers. Manag. 105, 617–629 (2015).

Article CAS Google Scholar

Gebremariam, S. N. & Marchetti, J. M. Biodiesel production through sulfuric acid catalyzed transesterification of acidic oil: Techno economic feasibility of different process alternatives. Energy Convers. Manag. 174, 639–648 (2018).

Article CAS Google Scholar

Miceli, M., Frontera, P., Macario, A. & Malara, A. Recovery/reuse of heterogeneous supported spent catalysts. Catalysts 11, 591 (2021).

Article CAS Google Scholar

Chellapandi, T. & Madhumitha, G. Montmorillonite clay-based heterogenous catalyst for the synthesis of nitrogen heterocycle organic moieties: A review. Mol. Divers. 1, 1–29 (2021).

Google Scholar

Faruque, M. O., Razzak, S. A. & Hossain, M. M. Application of heterogeneous catalysts for biodiesel production from microalgal oil: A review. Catalysts 10, 1025 (2020).

Article CAS Google Scholar

Kiss, A. A., Omota, F., Dimian, A. C. & Rothenberg, G. The heterogeneous advantage: Biodiesel by catalytic reactive distillation. Top Catal. 40, 141–150 (2006).

Article CAS Google Scholar

Fadhel, A. Z., Pollet, P., Liotta, C. L. & Eckert, C. A. Combining the benefits of homogeneous and heterogeneous catalysis with tunable solvents and nearcritical water. Molecules 15, 8400–8424 (2010).

Article CAS PubMed PubMed Central Google Scholar

Trindade, A. F., Gois, P. M. & Afonso, C. A. Recyclable stereoselective catalysts. Chem. Rev. 109, 418–514 (2009).

Article CAS PubMed Google Scholar

Ling, J. S. J. et al. A review of heterogeneous calcium oxide based catalyst from waste for biodiesel synthesis. SN Appl. Sci. 1, 1–8 (2019).

Article CAS Google Scholar

Lunardi, V. B. et al. Efficient one-step conversion of a low-grade vegetable oil to biodiesel over a zinc carboxylate metal-organic framework. ACS Omega 6, 1834–1845 (2021).

Article CAS PubMed PubMed Central Google Scholar

Haber, J. Molybdenum compounds in heterogeneous catalysis. In Molybdenum: An Outline of its Chemistry and Uses (eds Braithwaite, E. R. & Haber, J.) 477–617 (Elsevier, 1994).

Chapter Google Scholar

Nakagaki, S. et al. Use of anhydrous sodium molybdate as an efficient heterogeneous catalyst for soybean oil methanolysis. Appl. Catal. A. 351, 267–274 (2008).

Article CAS Google Scholar

Ferreira Pinto, B. et al. Effect of calcination temperature on the application of molybdenum trioxide acid catalyst: Screening of substrates for biodiesel production. Fuel 239, 290–296 (2019).

Article CAS Google Scholar

Mohebbi, S., Rostamizadeh, M. & Kahforoushan, D. Effect of molybdenum promoter on performance of high silica MoO3/B-ZSM-5 nanocatalyst in biodiesel production. Fuel 266, 117063 (2020).

Article CAS Google Scholar

Sankaranarayanan, T. M., Pandurangan, A., Banu, M. & Sivansanker, S. Transesterification of sunflower oil over MoO3 supported on alumina. Appl. Catal. A. 409, 239–247 (2011).

Article CAS Google Scholar

Bail, A. et al. Investigation of a molybdenum-containing silica catalyst synthesized by the sol-gel process in heterogeneous catalytic esterification reactions using methanol and ethanol. Appl. Catal. B. 130, 314–324 (2013).

Article CAS Google Scholar

Sankaranarayanan, T. M., Thirunavukkarasu, K., Banu, M., Pandurangan, K. & Sivansanker, S. Activity of supported MoO3 catalysts for the transesterification of sunflower oil. Int. J. Adv. Eng. Sci. Appl. Math. 5, 197–209 (2013).

Article Google Scholar

Mouat, A. R. et al. Reactivity of a carbon-supported single-site molybdenum dioxo catalyst for biodiesel synthesis. ACS Catal. 6, 6762–6769 (2016).

Article CAS Google Scholar

Navajas, A. et al. Catalytic performance of bulk and Al2O3-supported molybdenum oxide for the production of biodiesel from oil with high free fatty acids content. Catalysts 10, 158 (2020).

Article CAS Google Scholar

Feng, L., Wang, K. Y., Powell, J. & Zhou, H. C. Controllable synthesis of metal-organic frameworks and their hierarchical assemblies. Matter. 1, 801–824 (2019).

Article Google Scholar

Düren, T., Millange, F., Férey, G., Walton, K. S. & Snurr, R. Q. Calculating geometric surface areas as a characterization tool for metal-organic frameworks. J. Phys. Chem. C. 111, 15350–15356 (2007).

Article CAS Google Scholar

Farha, O. K. et al. Designing higher surface area metal-organic frameworks: Are triple bonds better than phenyls. J. Am. Chem. Soc. 134, 9860–9863 (2012).

Article CAS PubMed Google Scholar

Healy, C. et al. The thermal stability of metal-organic frameworks. Coord. Chem. Rev. 419, 213388 (2020).

Article CAS Google Scholar

Kaur, J. & Kaur, G. Review on flexible metal-organic frameworks. ChemistrySelect 6, 8227–8243 (2021).

Article CAS Google Scholar

Ghorbani-Choghamarani, A., Taherinia, Z. & Mohammadi, M. Facile synthesis of Fe3O4@ GlcA@ Ni-MOF composites as environmentally green catalyst in organic reactions. Environ. Technol. Innov. 24, 102050 (2021).

Article CAS Google Scholar

Ghorbani-Choghamarani, A. & Taherinia, Z. Fe3O4@ GlcA@ Cu-MOF: A magnetic metal-organic framework as a recoverable catalyst for the hydration of nitriles and reduction of isothiocyanates, isocyanates, and isocyanides. ACS Comb Sci. 22, 902–909 (2020).

Article CAS PubMed Google Scholar

Ghorbani-Choghamarani, A. & Taherinia, Z. Chiral cobalt-peptide metal-organic framework (Co-P-MOF): As an efficient and reusable heterogeneous catalyst for the asymmetric sulfoxidative cross-coupling reaction using poly sulfinylpiperazine. Synth. Met. 263, 116362 (2020).

Article CAS Google Scholar

Zhao, X., Wang, Y., Li, D. S., Bu, X. & Feng, P. Metal-organic frameworks for separation. Adv. Mater. 30, 1705189 (2018).

Article CAS Google Scholar

Du, J. et al. HKUST-1 MOFs decorated 3D copper foam with superhydrophobicity/superoleophilicity for durable oil/water separation. Colloids Surf. A. 573, 222–229 (2019).

Article CAS Google Scholar

Li, H. et al. Porous metal-organic frameworks for gas storage and separation: Status and challenges. J. Energy Chem. 1, 100006 (2019).

Google Scholar

Qazvini, O. T., Babarao, R. & Telfer, S. G. Selective capture of carbon dioxide from hydrocarbons using a metal-organic framework. Nat. Commun. 12, 1–8 (2021).

Article CAS Google Scholar

Abou-Elyazed, A. S., Ye, G., Sun, Y. & El-Nahas, A. M. A series of UiO-66(Zr)-structured materials with defects as heterogeneous catalysts for biodiesel production. Ind. Eng. Chem. Res. 58, 21961–21971 (2019).

Article CAS Google Scholar

Hassan, H. M. A., Betiha, M. A., Mohamed, S. K., ElSharkawy, E. A. & Ahmed, E. A. Salen-Zr(IV) complex grafted into amine-tagged MIL-101(Cr) as a robust multifunctional catalyst for biodiesel production and organic transformation reactions. Appl. Surf. Sci. 412, 394–404 (2017).

Article ADS CAS Google Scholar

Xie, W. & Wan, F. Immobilization of polyoxometalate-based sulfonated ionic liquids on UiO-66-2COOH metal-organic frameworks for biodiesel production via one-pot transesterification-esterification of acidic vegetable oils. Chem. Eng. J. 365, 40–50 (2019).

Article CAS Google Scholar

Xie, W. & Wan, F. Basic ionic liquid functionalized magnetically responsive Fe3O4@ HKUST-1 composites used for biodiesel production. Fuel 220, 248–256 (2018).

Article CAS Google Scholar

Xie, W. & Wang, H. Grafting copolymerization of dual acidic ionic liquid on core-shell structured magnetic silica: A magnetically recyclable Brönsted acid catalyst for biodiesel production by one-pot transformation of low-quality oils. Fuel 283, 118893 (2021).

Article CAS Google Scholar

Xie, W. & Wang, H. Immobilized polymeric sulfonated ionic liquid on core-shell structured Fe3O4/SiO2 composites: A magnetically recyclable catalyst for simultaneous transesterification and esterifications of low-cost oils to biodiesel. Renew. Energy. 145, 1709–1719 (2020).

Article CAS Google Scholar

Shokrani, R. & Haghighi, M. Textural evolution of hierarchical nanostructured ZSM-5 via sono-hydrothermal design by various carbon shapes for efficient biodiesel production. Appl. Catal. B. 271, 118940 (2020).

Article CAS Google Scholar

Lee, H. V., Juan, J. C. & Taufiq-Yap, Y. H. Preparation and application of binary acid–base CaO-La2O3 catalyst for biodiesel production. Renew. Energy. 74, 124–132 (2015).

Article CAS Google Scholar

Zhang, H. et al. Batch and continuous esterification for the direct synthesis of high qualified biodiesel from waste cooking oils (WCO) with Amberlyst-15/Poly (vinyl alcohol) membrane as a bifunctional catalyst. Chem. Eng. J. 388, 124214 (2020).

Article CAS Google Scholar

Wang, Y.-T. et al. One-step production of biodiesel from Jatropha oils with high acid value at low temperature by magnetic acid-base amphoteric nanoparticles. Chem. Eng. J. 348, 929–939 (2018).

Article CAS Google Scholar

Zhang, F., Li, X., Yang, L., Zhang, Y. & Zhang, M. A Mo-based metal-organic framework toward improving flame retardancy and smoke suppression of epoxy resin. Polym. Adv. Technol. 32, 3266–3277 (2021).

Article CAS Google Scholar

Zhang, X., Li, D., Dong, C., Shi, J. & Xu, Y. The synergistic supercapacitive performance of Mo–MOF/PANI and its electrochemical impedance spectroscopy investigation. Mater. Today Commun. 21, 100711 (2019).

Article CAS Google Scholar

Martían-Zarza, P., Arrieta, J. M., Muñoz-Roca, M. C. & Gili, P. Synthesis and characterization of new octamolybdates containing imidazole, 1-methyl-or 2-methyl-imidazole co-ordinatively bound to molybdenum. J. Chem. Soc. Dalton trans. 10, 1551 (1993).

Article Google Scholar

Du, J., Chai, J., Li, Q., Zhang, W. & Tang, B. Application of two-dimensional layered Mo–MOF@ ppy with high valency molybdenum in lithium-ion batteries. Colloids Surf. A Physicochem. Eng. Asp. 632, 127810 (2022).

Article CAS Google Scholar

Xu, J., Ji, Q., Yan, X., Wang, C. & Wang, L. Ni (acac)2/Mo–MOF-derived difunctional MoNi@ MoO2 cocatalyst to enhance the photocatalytic H2 evolution activity of g-C3N4. Appl. Catal. B. 268, 118739 (2020).

Article CAS Google Scholar

Yao, M. S. et al. Gas transport regulation in a MO/MOF interface for enhanced selective gas detection. J. Mater. Chem. A. 7, 18397–18403 (2019).

Article CAS Google Scholar

Zhang, Y., Wong, W. T. & Yung, K. F. Biodiesel production via esterification of oleic acid catalyzed by chlorosulfonic acid modified zirconia. Appl. Energy. 116, 191–198 (2014).

Article ADS CAS Google Scholar

Shu, Q. et al. Preparation of the F−SO42−/MWCNTs catalyst and kinetic studies of the biodiesel production via esterification reaction of oleic acid and methanol. Renew. Energy. 135, 836–845 (2019).

Article CAS Google Scholar

Li, Y., Hu, S., Cheng, J. & Lou, W. Acidic ionic liquid-catalyzed esterification of oleic acid for biodiesel synthesis. Chin. J. Catal. 35, 396–406 (2014).

Article CAS Google Scholar

Wang, Y. T. et al. Biodiesel production from esterification of oleic acid by a sulfonated magnetic solid acid catalyst. Renew. Energy. 139, 688–695 (2019).

Article CAS Google Scholar

Gong, S. W., Lu, J., Wang, H. H., Liu, L. J. & Zhang, Q. Biodiesel production via esterification of oleic acid catalyzed by picolinic acid modified 12-tungstophosphoric acid. Appl. Energy. 134, 283–289 (2014).

Article CAS Google Scholar

Download references

The authors would like to thank the research facilities of Bu-Ali Sina University and Ilam University for the financial support of this research project.

Department of Organic Chemistry, Faculty of Chemistry, Bu-Ali Sina University, Hamedan, 6517838683, Iran

Arash Ghorbani-Choghamarani

Department of Chemistry, Faculty of Science, Ilam University, Ilam, Iran

Zahra Taherinia & Yunes Abbasi Tyula

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

A.G.: Choghamarani: Funding acquisition, Supervision, Project administration, Conceptualization, Resources, Writing-review & editing. Z.T.: Methodology, Validation, Investigation. Y.A.: Writing—original draft.

Correspondence to Arash Ghorbani-Choghamarani.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Ghorbani-Choghamarani, A., Taherinia, Z. & Tyula, Y.A. Efficient biodiesel production from oleic and palmitic acid using a novel molybdenum metal–organic framework as efficient and reusable catalyst. Sci Rep 12, 10338 (2022). https://doi.org/10.1038/s41598-022-14341-4

Download citation

Received: 20 March 2022

Accepted: 06 June 2022

Published: 20 June 2022

DOI: https://doi.org/10.1038/s41598-022-14341-4

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.