Next Article in Journal
Toward Evaluating Critical Factors of Extubation Outcome with XCSR-Generated Rules
Next Article in Special Issue
Kinetics of Biodiesel Production from Microalgae Using Microbubble Interfacial Technology
Previous Article in Journal
The Allogenic Dental Pulp Transplantation from Son/Daughter to Mother/Father: A Follow-Up of Three Clinical Cases
Previous Article in Special Issue
Deep Eutectic Solvents for Biodiesel Purification in a Microextractor: Solvent Preparation, Selection and Process Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Zr-Based Metal-Organic Frameworks for Green Biodiesel Synthesis: A Minireview

1
College Rural Revitalization Research Center of Guizhou, Anshun University, Anshun 561000, China
2
School of Chemistry and Chemical Engineering, Anshun University, Anshun 561000, China
3
School of Resource and Environmental Engineering, Anshun University, Anshun 561000, China
*
Authors to whom correspondence should be addressed.
Bioengineering 2022, 9(11), 700; https://doi.org/10.3390/bioengineering9110700
Submission received: 12 October 2022 / Revised: 6 November 2022 / Accepted: 15 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue Acceleration of Biodiesel Production)

Abstract

:
Metal–organic frameworks (MOFs) have widespread application prospects in the field of catalysis owing to their functionally adjustable metal sites and adjustable structure. In this minireview, we summarize the current advancements in zirconium-based metal–organic framework (Zr-based MOF) catalysts (including single Zr-based MOFs, modified Zr-based MOFs, and Zr-based MOF derivatives) for green biofuel synthesis. Additionally, the yields, conversions, and reusability of Zr-based MOF catalysts for the production of biodiesel are compared. Finally, the challenges and future prospects regarding Zr-based MOFs and their derivatives for catalytic application in the biorefinery field are highlighted.

1. Introduction

With rapid population growth and high industrial energy demand from fossil fuel resources, fossil fuel usage leading to the energy security crisis and climate change (e.g., greenhouse gas emissions) has been of great concern [1]. Based on this, it is necessary to search for alternative fuel resources. Currently, renewable biofuels derived from biomass have gained enormous attention, and one of these liquid biofuels is biodiesel [2]. Chemically, biodiesel (fatty acid alkyl ester, FAME) is generally produced via the transesterification of edible oils (e.g., rapeseed oil, palm oil, sunflower oil), non-edible oils (e.g., Jatropha curcus, Euphorbia lathyris L., Monotheca buxifolia, Sinapis Arvensis), microalgal oils, or waste cooking oil or via the esterification of free fatty acids (e.g., oleic acid, lauric acid, palmitic acid) and methanol or ethanol using acid/alkali as catalysts [3,4,5,6,7,8].
Traditionally, biodiesel production is carried out using liquid acid/alkali catalysts due to their high catalytic activity. Unfortunately, these homogeneous catalytic processes exhibit numerous disadvantages, such as high operating costs for steps such as product purification, catalyst neutralization, and a large amount of industrial wastewater that requires treatment [9]. With regard to this, the utilization of heterogeneous catalysts is becoming an efficient candidate for the production of biodiesel, because of their simple recovery, ease of reuse, insolubility in reaction solvents, and reduction in waste treatment [10]. According to Figure 1, various kinds of heterogeneous acid/alkali catalysts are available for different types of organic reactions, including metal oxides, ionic liquids, heteropoly acids, zeolites, sulfonic-acid-functionalized catalysts, etc. [11,12]. However, some problems such as leaching, lesser activity, low stability, and longer reaction time for some heterogeneous catalysts are found in the transesterification/esterification reaction process [13].
Quite recently, metal–organic frameworks (MOFs) have gained enormous attention due to their unique features, such as great specific surface area, uniformity in pore size, large porosity, adjustable properties, tunable structures, and controllable functional groups [14]. In addition, MOFs are suitable to be functionalized by coordinating acid/base functional groups, which have been widely studied in catalysis [15,16,17]. Among the numerous types of MOFs, zirconium-based MOFs (Zr-based MOFs) have been frequently applied as potential porous materials owing to the presence of Lewis and Brønsted acidity [18,19]. At present, several reviews have summarized the applications of Zr-MOFs in catalysis [20,21]. However, none has given a detailed study on the applications of Zr-MOFs and derivatives for green biodiesel synthesis. Thus, the present review focuses on the current development of Zr-based MOFs and derivatives for green biofuel production. More importantly, the catalytic performance and reusability of single Zr-based MOFs, modified Zr-based MOFs, and Zr-based MOF derivatives are systematically discussed. Finally, the conclusions and prospects are emphasized.

2. Zr-Based MOF Catalysts

In 2008, a zirconium-based inorganic building brick (Zr-MOFs) was first reported by Lillerud et al. [22], and the results showed that Zr-MOFs possess high stability due to the combination of strong Zr-O bonds, the inner Zr6 cluster, and the addition of μ3-OH groups. A large number of Zr-based MOF catalysts have been widely used for fuel synthesis, including UiO-66 (see Figure 2), UiO-67, MOF-801, MOF-808, UiO-66-NH2, etc., owing to their excellent chemical and thermal stability under harsh conditions, large specific surface areas, smaller particle size, and strong acid sites by the tuning of structural defects [23,24].

2.1. Single Zr-Based MOF Catalysts

Single Zr-based MOFs usually show fewer Lewis acid characteristics due to the saturated Zr atom, and they have exhibited less catalytic activity in acid-catalyzed organic reactions. In view of this, the synthesis of various zirconium-containing UiO-66 samples by varying the synthesis temperatures and terephthalic acid/ZrCl4 ratios was reported by Zhou, et al. [25]. They found that UiO-66 catalysts with different amounts of defects could be synthesized under various synthesis conditions, and the catalytic activities of the UiO-66 depended on the defect amount. The obtained catalyst was used to catalyze the transesterification of soybean oil with methanol, and a conversion rate of 98.5% was acquired through a catalyst amount of 9%, with an oil/methanol molar ratio of 1:40, and at 140 °C for 5 h. UiO-66 with defects can be especially easily reused. Caratelli and his co-workers [26] also utilized UiO-66 MOFs as an acid catalyst for the production of ethyl levulinate via esterification from levulinic acid and ethanol. Various defective hydrated and dehydrated UiO-66 materials demonstrated excellent performance, and a maximum yield (>70%) of ethyl levulinate was achieved.
A similar study was conducted by Jrad, et al. [27], in which three isostructural Zr-based MOFs (UiO-66, UiO-66(COOH)2, and UiO-66(NH2)) were prepared (Figure 3). UiO-66(COOH)2 demonstrated superior catalytic activity in the esterification of butyric acid and butanol, and a 90% conversion rate of butyl butyrate was achieved. This better catalytic activity was possibly related to the smaller particle size of the catalyst and the additional active acid functional groups grafted onto the original organic linker.
Wei’s group [28] also designed a series of defective UiO-66 catalysts for the esterification of levulinic acid with ethanol. The results showed that the synergistic effects between unsaturated Zr6 nodes and hydroxyl groups can have a significant influence on catalytic activity. The synthesized defective UiO-66 catalyst possessed excellent stability and could retain 75% of its initial activity after five cycles. Similarly, Chaemchuen and co-workers [29] synthesized UiO-66 catalyst for the esterification reaction between oleic acid and methanol. According to the kinetic analysis, an activation energy of 54.9 ± 1.8 kJ/mol was obtained, and the desorption of methyl oleate was found to be irreversible. Desidery, et al. [30] tested MOF-808 for the conversion of dimethyl carbonate into ethyl methyl carbonate. The MOF-808 catalyst exhibited superior catalytic performance and could be recycled for up to four cycles without any major change in activity or structure. Shaik, et al. [31] developed the Zr-fumarate MOF (MOF-801) as a heterogeneous catalyst for biodiesel production. The characterization results demonstrated that MOF-801 possessed cubic structure, high crystallinity, good thermal stability, and moderate catalytic activity. Under optimal reaction conditions, the conversion rate of used vegetable oil was 60%; the activity of MOF-801 is probably due to the cationic Zr and anionic O2 sites in the crystal structure.
Besides this, de la Flor, et al. [32] reported the synthesis of defective UiO-66(Zr) catalyst for the production of jet-fuel precursors via aldol-condensation, and total furfural conversion and selectivity (~100%) were obtained. Rapeyko, et al. [33] also reported that the as-synthesized UiO-66 could efficiently catalyze the selective ketalization of levulinic acid and 1,2-propanediol, and high selectivity (91–93%) was attained.
From the studied literature, it is derived that the single Zr-based MOFs can be considered as a catalyst for acid-catalyzed reactions. However, the activity of single Zr-based MOFs still needs to be further improved, and facilely tuning the defect density on nodes by introducing modulators may be a very interesting approach to developing highly active Zr-based MOF catalysts.

2.2. Modified Zr-Based MOF Catalysts

To improve the catalytic activity and chemical stability of Zr-based MOFs, several researchers have investigated various types of modification methods, including functional organic linkers, loading active components (e.g., lipase, ionic liquids, heteropoly acid, etc.), and the incorporation of metal ions into Zr-based MOFs (Table 1).
Table 1. Recent findings on green fuel production using modified Zr-based MOF catalysts.
Table 1. Recent findings on green fuel production using modified Zr-based MOF catalysts.
EntryRaw MaterialCatalystReaction Conditions (Time, Temperature, Catalyst Amount, Molar Ratio (Acid(Oil):Alcohol))Yield (Y/%) or Conversion (C/%)ReusabilityEa (KJ/mol)Ref.
1Lauric acid + MethanolUiO-66-NH22 h, 60 °C, 8%, 1:26Y > 99Not reported\[34]
2Levulinic acid + n-butanolUiO-66-NH25 h, 120 °C, 1.8%, 1:6Y = 99%3 cycles, no
significant loss
\[35]
3Levulinic acid + EthanolUiO-66-(COOH)224 h, 78 °C, 0.39%, 1:20Y = 97%5 cycles, Y = 93.9%\[36]
4Oleic acid + MethanolUiO-66(Zr)-NH24 h, 60 °C, 6%, 1:39C = 97%4 cycles, C > 50%15.13[37]
5Oleic acid + Methanol 10SA/UiO-66(Zr)4 h, 25 °C, 6%, 1:39C = 94.5%6 cycles, C = 83%32.53[38]
6Levulinic acid + EthanolUiO66-SO3H(100)6 h, 80 °C, 0.4%, 1:10Y = 87%4 cycles, Y = 84%\[39]
7Ricinus communis oil + Methanol Lipase/Zr-MOF/PVP12 h, 50 °C, 2 mg, 1:3C = 83%7 cycles, C = 66%\[40]
8Oleic acid + MethanolUiO-G2 h, 70 °C, 8%, 1:12C = 91.3%4 cycles, C = 66.6%28.61[41]
9Acetic acid + Isooctyl alcoholUiO-67-CF3SO318 h, 90 °C, 0.2 g, 6:1C = 98.6%5 cycles, C = 95.9%\[42]
10Tripalmitin + MethanolUiO-66-[C3NH2] [SO3CF3]12 h, 85 °C, 0.025 g, 1:121.5Y = 86.6–98.4%Not reported38.9[43]
11Jatropha oil + MethanolPSH/UiO-66-NO24 h, 70 °C, 4%, 1:25C= 97.57%3 cycles, C= 77.14%\[44]
12Oleic acid + Methanol AIL@NH2-UiO-666 h, 75 °C, 5%, 1:14C = 95.22%6 cycles, C = 90.42%\[45]
13Oleic acid + MethanolCa2+/UiO-66(Zr)4 h, 60 °C, 6%, 1:39Y = 98%5 cycles, Y = 84%36.73[46]
14Oleic acid + MethanolK-PW12@UIO-66(Zr)4 h, 75 °C, 5%, 1:20C = 90%10 cycles, no
significant loss
\[47]
15Acetic acid + n-butanolHPW@UiO-663 h, 120 °C, 3%, 1:2C = 80.2%4 cycles, C = 63%\[48]
16Soybean oil + C8 + C10Cs2.5H0.5PW12O40@UiO-6610 h, 150 °C, 7%,1:5:5FA incorporation =20.3%5 cycles, no
significant loss
\[49]
17Soybean oil + Methanol AILs/HPW/UiO-66-2COOH6 h, 110 °C, 10%, 1:35C = 95.8%5 cycles, C > 80%\[50]
18Euphorbia
Lathyris L. oil + Methanol
HPW/UiO-66-NH28 h, 180 °C, 3.5%, 1:40Y = 91.2%4 cycles, no
significant loss
31.0[51]
19Oleic acid + MethanolFDCA/SA-UiO-66(Zr)24 h, 60 °C, 6%, 1:40Y = 98.4%6 cycles, Y > 90%\[52]
20Soybean oil + MethanolPW12@UIO-664 h, 75 °C, 0.2 g, 1 g:5.5 mlC = 91.1%4 cycles, no
significant loss
\[53]
21Lauric acid + MethanolHSiW-UiO-664 h, 160 °C, 7%, 1:20C = 80.5%4 cycles, C = 70.2%27.5[54]
22Oleic acid + MethanolZrSiW/UiO-664 h, 150 °C, 8%, 1:20C = 98.0%4 cycles, C = 88.9%\[55]
23Lauric acid + MethanolAg1(NH4)2PW12O40/UiO-663 h, 150 °C, 10%, 1:15C = 75.6%4 cycles, C = 70.6%35.2[56]
24Oleic acid + MethanolCe-BDC@HSiW@UiO-664 h, 130 °C, 0.2 g, 1:30C = 81.5%6 cycles, C = 76.9%\[57]
Cirujano’s group [34,35] prepared an UiO-66-NH2 catalyst to convert lauric acid to methyl laurate via an esterification reaction. The high activity of UiO-66-NH2 with respect to UiO-66 is attributed to the occurrence of cooperative acid–base catalysis in the frame network. In addition, UiO-66-NH2 has been successfully used for esterification of levulinic acid with various alcohols. A possible bifunctional acid–base catalyst mechanism for esterification was proposed, as displayed in Figure 4.
Likewise, Wang, et al. [36] employed UiO-66-(COOH)2 as a heterogeneous catalyst for the esterification of levulinic acid, and Abou-Elyazed, et al. [37] also employed UiO-66(Zr)-NH2 for the esterification of oleic acid. Meanwhile, Abou-Elyazed’s group [38] also demonstrated the direct preparation of Ca2+-doped UiO-66(Zr) under solvent-free conditions. In detail, the introduction of Ca2+ could greatly enhance the catalytic performance and stability in the esterification because of the existence of double active sites with the formation of more defects.
Desidery, et al. [39] investigated partially and fully sulfonated hydrated UiO66 catalysts prepared by one-step solvothermal synthesis. Compared to that of commercial Amberlyst 15, the activity of the fully sulfonated hydrated UiO66 afforded the highest yield of ethyl levulinate.
UiO-66(Zr) used as a support for p-toluenesulfonic acid (PTSA) through a defect coordination strategy was proposed by Li, et al. [41]. Their results indicated that the PTSA was successfully introduced into UiO-66(Zr), and the highest conversion rate of oleic acid to biodiesel of 91.3% was acquired under mild conditions. More specifically, a reusability study showed that the conversion was dramatically reduced from 91.3% to 76.65% after four cycles, and they verified a loss of Zr and S in the reaction system.
Recently, acidic or basic ionic liquids (ILs) have shown efficient catalytic activities in various organic reactions. However, they also suffer from several shortcomings, such as high viscosity, diffusion limitations, and difficulty in separation. In order to overcome these problems, the introduction of ILs into Zr-based MOFs has been studied [42,43,44,45]. Acidic ILs (AIL) were introduced into the NH2-UiO-66 matrix (See Figure 5) via acid–base interaction by Lu, et al. [45]. Accordingly, the best mass ratio of AIL to NH2-UiO-66 in 3AIL/NH2-UiO-66 displayed excellent activity and reusability in the esterification of oleic acid; a conversion rate of 95.22% was achieved in 6 h, and it could still reach 90.42% conversion after six cycles. Moreover, it was concluded that the good conversion rate was attributed to the stimulating synergy between the -SO3H group of AIL and the -NH2 group of NH2-UiO-66 on the MOFs.
Apart from active ILs, heteropoly acids (HPAs) with structural diversity and tunable Brønsted/Lewis acidity have also been reported as efficient acid catalysts for biodiesel synthesis. This is despite their high activity, solubility in many polar solvents, and very low surface area, which limit their application for catalysis. Therefore, the loading of various HPAs on Zr-based MOF materials has been performed [47,48].
Xie’s group [50] studied the one-pot transesterification–esterification of acidic vegetable oils to produce biodiesel by employing UiO-66-2COOH modified with HPW and sulfonated ILs as an acid catalyst (AILs/POM/UiO-66-2COOH). In their study, the surface area, pore volume, and mean pore size of the as-prepared composite catalyst were found to be 8.63 m2/g, 0.04 cm3/g, and 16.07 nm, respectively. Furthermore, the highest observed conversion rate was 95.8%, and the solid catalyst could maintain high activity even when 9 wt% free fatty acid and 3 wt% water were added into the feedstock.
As reported in much of the literature, Yang’s group [51] synthesized HPW/UiO-66-NH2 Lewis/Brønsted acid bifunctional hybrid catalyst by the electrovalent assembly of HPW and UiO-66-NH2. The resulting HPW/UiO-66-NH2 exhibited a highest acid density of 1.7 mmol/g, larger surface area of 301.6 m2/g, and both Lewis and Brønsted acid sites. The biodiesel yield obtained from the (trans)esterification of Lathyris L. oil was more than 91.2%. Notably, the composite catalyst was reusable for four cycles with no significant decrease in activity, and hot filtration experiments showed that the composite has heterogeneous characteristics. Another study conducted by Yang’s group [52] examined the synthesis of FDCA/SA-UiO-66(Zr) catalyst by a facile grinding method. Accordingly, DCA/SA-UiO-66(Zr) demonstrated superior or equivalent catalytic activity in the esterification of oleic acid due to its Lewis acidity and hydrophobicity.
Recently, our group also studied the production of biodiesel from free fatty acid with methanol over HPAs or doped HPAs incorporated into UiO-66 frameworks (e.g., ZrSiW/UiO-66, Ag1(NH4)2PW12O40/UiO-66, and Ce-BDC@HSiW@UiO-66) [54,55,56,57]. All these composite catalysts exhibited good catalytic activity and reusability. Table 1 summarizes the modified Zr-based MOF catalysts used for biofuel production. As can be seen here, many researchers agree that modified Zr-based MOF catalysts can effectively catalyze esterification or transesterification processes.

2.3. Zr-Based MOF-Derived Catalysts

Recently, MOFs have also been employed as a precursor substrate and template support for derived material synthesis. The synthesis of porous carbon and metal oxide via a thermal decomposition process was first reported by Xu’s group [58,59]. Since then, MOF derivatives have been attracted increasing attention as novel catalysts. In particular, the pyrolysis of defective Zr-based MOFs can provide a promising platform for various functional materials’ synthesis.
Lu, et al. [60] successfully synthesized flower-like mesoporous sulfated zirconia nanosheets via the thermal decomposition of in situ sulfated Zr-MOFs as the S/Zr ratio increased to 0.5. Investigations on the sulfated zirconia nanosheets at a calcination temperature of 500 °C showed a large surface area (186.1 m2/g) and strong interaction between the sulfate and zirconia atoms, affording excellent catalytic performance and stability for the production of biodiesel. Besides this, the mechanism of transesterification was studied, as shown in Figure 6.
Li, et al. [61] employed a UiO-66(Zr) support impregnated with calcium acetate for CaO/ZrO2 catalyst synthesis via an activation process in nitrogen (UCN) and air (UCA) atmosphere. Among these catalysts, UCN650 calcined at 650 °C attained a relatively large specific surface area (24.06 m2/g); meanwhile, the catalyst generated active sites of CaxZryOx+2y and CaO inside and was shown to be effective in catalyzing palm oil transesterification, reaching a maximum conversion rate of 98.2%. Moreover, the UCN650 catalyst maintained its catalytic properties when it was recycled three times. The properties of the resulting biodiesel (density, kinetic viscosity, acid value, etc.) were also found to comply with the EN 14214 standards.
Our group also employed UiO-66 as a precursor for HSiW@ZrO2 hybrid synthesis, and SEM images of HSiW@UiO-66 at different calcination temperatures (300 °C, 400 °C, 500 °C) are shown in Figure 7. Nanoporous HSiW@ZrO2 was obtained by calcinating at 300 °C, exhibiting relatively high surface area (338 m2/g), appropriate pore size (2.5 nm), strong acidity (6.2 mmol/g), and the highest catalytic activity in the esterification of oleic acid; its conversion rate was high at 94.0% and stayed above 80% after nine catalytic cycles [62].
Dimethyl ether (DME) has gained attention for its application as a second-generation fuel, and it can be synthesized through a methanol dehydration process. Goda, et al. [63] used UiO-66 as a precursor for the synthesis of ZrOSO4@C catalyst. In the experiment, it was observed that ZrOSO4@C has weak and intermediate acidic sites and could be effectively applied for methanol dehydration to DME, with the highest conversion (100%) and selectivity (100%).
Hong, et al. [64] prepared 3D porous Cu@ZrOx catalysts via in situ reconstruction of size-confined Cu@UiO-66 for methanol synthesis from CO2 hydrogenation, and the optimized catalyst exhibited quite high methanol selectivity of 78.8% at 260 °C and 4.5 MPa, attributed to the many Cu+−ZrOx interfaces present as active sites in the material framework. Zeng’s group [65] also designed and prepared porous hydrous zirconia via UiO-66 pyrolysis as a support for NiII centers. In methane production from CO2 hydrogenation, the resultant catalyst exhibited excellent activity and stability.
Based on the literature, Zr-based MOF-derived materials with stable porous structures and many active sites are expected to be widely used for the development of high-performance composite catalysts in the future.

3. Conclusions and Future Prospects

Herein, a comprehensive review was presented on the synthesis of Zr-based MOFs and their derived composite materials for their catalytic application in green biofuel synthesis in recent years. The current review attempted to thoroughly demonstrate the use of single Zr-based MOF catalysts, modified Zr-based MOF catalysts, and Zr-based MOF-derived catalysts in the literature. Their high surface area, adjustable pore structure, acceptable recyclability, and strong acid sites obtained by tuning structural defects make Zr-based MOFs suitable for esterification or transesterification process.
However, looking ahead, many challenges for the large-scale application of Zr-based MOFs still exist, such as the design and development of inexpensive Zr-based MOFs at an industrial scale with high yields. The self-assembly mechanism of Zr-based MOFs and their derived materials is still unclear, and further exploration via both experimental and theoretical approaches is still required. The chemical and thermal stability of Zr-based MOFs is still not adequate and needs to be further improved. Facilely tuning the defect density on nodes by exploring new modification approaches could bring beneficial changes to the catalytic performance. By combining Zr-based MOFs with appropriate active materials such as enzymes, graphene derivatives, and magnetic substances, composite materials could be synthesized to improve their catalytic performance. Controlling the structure, composition, and distribution of the active component of Zr-based MOF-derived catalysts, while aiming to maintain the original structure of the Zr-based MOFs, still needs to be further studied.
As a whole, the application of Zr-based MOFs and catalysts derived from them is important not only for green biodiesel synthesis but also for the conversion of biomass. Despite facing many challenges, hopefully, the existing issues will be resolved sooner or later, and the application prospects of biorefineries will also be very bright.

Author Contributions

Q.Z. and J.W. jointly conceived the manuscript and discussed the outline; Q.Z. wrote the manuscript; S.Z., J.M. and J.C. made preliminary revisions to the manuscript; Q.Z. and Y.Z. were in charge of project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22262001), the Guizhou Science and Technology Foundation ((2020)1Y054, 20181401), the 2018 Thousand Level Innovative Talents Training Program of Guizhou Province, the Project of Anshun University supporting Doctors Research ((2021)asxybsjj01), the Creative Research Groups Support Program of Guizhou Education Department (KY (2017)049), and the Youth Growth S&T Personnel Foundation of the Guizhou Education Department (KY (2018)321).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare there is no conflict of interest regarding the publication of this paper.

References

  1. Hoang, A.T.; Tabatabaei, M.; Aghbashlo, M.; Carlucci, A.P.; Ölçer, A.I.; Le, A.T.; Ghassemi, A. Rice bran oil-based biodiesel as a promising renewable fuel alternative to petrodiesel: A review. Renew. Sustain. Energy Rev. 2020, 135, 110204. [Google Scholar] [CrossRef]
  2. Bekhradinassab, E.; Tavakoli, A.; Haghighi, M.; Shabani, M. Catalytic biofuel production over 3D macro-structured cheese-like Mn-promoted TiO2 isotype: Mn-catalyzed microwave-combustion design. Energy Convers. Manag. 2021, 251, 114916. [Google Scholar] [CrossRef]
  3. Rezania, S.; Mahdinia, S.; Oryani, B.; Cho, J.; E Kwon, E.; Bozorgian, A.; Nodeh, H.R.; Darajeh, N.; Mehranzamir, K. Biodiesel production from wild mustard (Sinapis Arvensis) seed oil using a novel heterogeneous catalyst of LaTiO3 nanoparticles. Fuel 2022, 307, 121759. [Google Scholar] [CrossRef]
  4. Zhang, Q.; Ling, D.; Lei, D.; Wang, J.; Liu, X.; Zhang, Y.; Ma, P. Green and Facile Synthesis of Metal-Organic Framework Cu-BTC-Supported Sn (II)-Substituted Keggin Heteropoly Composites as an Esterification Nanocatalyst for Biodiesel Production. Front. Chem. 2020, 8, 129. [Google Scholar] [CrossRef] [PubMed]
  5. Cholapandian, K.; Gurunathan, B.; Rajendran, N. Investigation of CaO nanocatalyst synthesized from Acalypha indica leaves and its application in biodiesel production using waste cooking oil. Fuel 2022, 312, 122958. [Google Scholar] [CrossRef]
  6. Rozina; Ahmad, M.; Elnaggar, A.Y.; Teong, L.K.; Sultana, S.; Zafar, M.; Munir, M.; Hussein, E.E.; Abidin, S.Z.U. Sustainable and eco-friendly synthesis of biodiesel from novel and non-edible seed oil of Monotheca buxifolia using green nano-catalyst of calcium oxide. Energy Convers. Manag. X 2022, 13, 100142. [Google Scholar] [CrossRef]
  7. Pan, H.; Xia, Q.; Li, H.; Wang, Y.; Shen, Z.; Wang, Y.; Li, L.; Li, X.; Xu, H.; Zhou, Z.; et al. Direct production of biodiesel from crude Euphorbia lathyris L. Oil catalyzed by multifunctional mesoporous composite materials. Fuel 2021, 309, 122172. [Google Scholar] [CrossRef]
  8. Pan, H.; Xia, Q.; Wang, Y.; Shen, Z.; Huang, H.; Ge, Z.; Li, X.; He, J.; Wang, X.; Li, L.; et al. Recent advances in biodiesel production using functional carbon materials as acid/base catalysts. Fuel Process. Technol. 2022, 237, 107421. [Google Scholar] [CrossRef]
  9. Zhang, Q.; Luo, Q.; Wu, Y.; Yu, R.; Cheng, J.; Zhang, Y. Construction of a Keggin heteropolyacid/Ni-MOF catalyst for esterification of fatty acids. RSC Adv. 2021, 11, 33416–33424. [Google Scholar] [CrossRef] [PubMed]
  10. Yan, W.; Zhang, D.; Sun, Y.; Zhou, Z.; Du, Y.; Du, Y.; Li, Y.; Liu, M.; Zhang, Y.; Shen, J.; et al. Structural sensitivity of heterogeneous catalysts for sustainable chemical synthesis of gluconic acid from glucose. Chin. J. Catal. 2020, 41, 1320–1336. [Google Scholar] [CrossRef]
  11. Chen, B.; Yan, G.; Chen, G.; Feng, Y.; Zeng, X.; Sun, Y.; Tang, X.; Lei, T.; Lin, L. Recent progress in the development of advanced biofuel 5-ethoxymethylfurfural. BMC Energy 2020, 2, 2. [Google Scholar] [CrossRef]
  12. Ji, J.; Bao, Y.; Liu, X.; Zhang, J.; Xing, M. Molybdenum-based heterogeneous catalysts for the control of environmental pollutants. EcoMat 2021, 3, e12155. [Google Scholar] [CrossRef]
  13. Sahar, J.; Farooq, M.; Ramli, A.; Naeem, A.; Khattak, N.S. Biodiesel production from Mazari palm (Nannorrhops ritchiana) seeds oil using Tungstophosphoric acid decorated SnO2@Mn-ZIF bifunctional heterogeneous catalyst. Appl. Catal. A Gen. 2022, 643, 118740. [Google Scholar] [CrossRef]
  14. Fang, R.; Dhakshinamoorthy, A.; Li, Y.; Garcia, H. Metal organic frameworks for biomass conversion. Chem. Soc. Rev. 2020, 49, 3638–3687. [Google Scholar] [CrossRef]
  15. Yadav, S.; Dixit, R.; Sharma, S.; Dutta, S.; Solanki, K.; Sharma, R.K. Magnetic metal–organic framework composites: Structurally advanced catalytic materials for organic transformations. Mater. Adv. 2021, 2, 2153–2187. [Google Scholar] [CrossRef]
  16. Zhang, Q.; Zhang, Y.; Cheng, J.; Li, H.; Ma, P. An Overview of Metal-organic Frameworks-based Acid/Base Catalysts for Biofuel Synthesis. Curr. Org. Chem. 2020, 24, 1876–1891. [Google Scholar] [CrossRef]
  17. Wei, Y.; Zhang, Y.; Li, B.; Guan, W.; Yan, C.; Li, X.; Yan, Y. Facile synthesis of metal-organic frameworks embedded in interconnected macroporous polymer as a dual acid-base bifunctional catalyst for efficient conversion of cellulose to 5-hydroxymethylfurfural. Chin. J. Chem. Eng. 2022, 44, 169–181. [Google Scholar] [CrossRef]
  18. Li, Y.; Meng, X.; Luo, R.; Zhou, H.; Lu, S.; Yu, S.; Bai, P.; Guo, X.; Lyu, J. Aluminum/Tin-doped UiO-66 as Lewis acid catalysts for enhanced glucose isomerization to fructose. Appl. Catal. A Gen. 2022, 632, 118501. [Google Scholar] [CrossRef]
  19. Ling, L.; Yang, W.; Yan, P.; Wang, M.; Jiang, H. Light-Assisted CO2 Hydrogenation over Pd3Cu@UiO-66 Promoted by Active Sites in Close Proximity. Angew. Chem. Int. Ed. 2022, 61, e202116396. [Google Scholar] [CrossRef] [PubMed]
  20. Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-based metal–organic frameworks: Design, synthesis, structure, and applications. Chem. Soc. Rev. 2016, 45, 2327–2367. [Google Scholar] [CrossRef] [PubMed]
  21. Dhakshinamoorthy, A.; Santiago-Portillo, A.; Asiri, A.M.; Garcia, H. Engineering UiO-66 Metal Organic Framework for Heterogeneous Catalysis. ChemCatChem 2019, 11, 899–923. [Google Scholar] [CrossRef]
  22. Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. [Google Scholar] [CrossRef] [PubMed]
  23. Schelling, M.; Kim, M.; Otal, E.; Hinestroza, J. Decoration of Cotton Fibers with a Water-Stable Metal-Organic Framework (UiO-66) for the Decomposition and Enhanced Adsorption of Micropollutants in Water. Bioengineering 2018, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  24. Cheng, J.; Qian, L.; Guo, H.; Mao, Y.; Shao, Y.; Yang, W. A new aminobenzoate-substituted s-triazin-based Zr metal organic frameworks as efficient catalyst for biodiesel production from microalgal lipids. Fuel Process. Technol. 2022, 238, 107487. [Google Scholar] [CrossRef]
  25. Zhou, F.; Lu, N.; Fan, B.; Wang, H.; Li, R. Zirconium-containing UiO-66 as an efficient and reusable catalyst for transesterification of triglyceride with methanol. J. Energy Chem. 2016, 25, 874–879. [Google Scholar] [CrossRef] [Green Version]
  26. Caratelli, C.; Hajek, J.; Cirujano, F.G.; Waroquier, M.; i Xamena, F.X.L.; Van Speybroeck, V. Nature of active sites on UiO-66 and beneficial influence of water in the catalysis of Fischer esterification. J. Catal. 2017, 352, 401–414. [Google Scholar] [CrossRef] [Green Version]
  27. Jrad, A.; Abu Tarboush, B.J.; Hmadeh, M.; Ahmad, M. Tuning acidity in zirconium-based metal organic frameworks catalysts for enhanced production of butyl butyrate. Appl. Catal. A Gen. 2019, 570, 31–41. [Google Scholar] [CrossRef]
  28. Wei, R.; Fan, J.; Qu, X.; Gao, L.; Wu, Y.; Zhang, Z.; Hu, F.; Xiao, G. Tuning the Catalytic Activity of UiO-66 via Modulated Synthesis: Esterification of Levulinic Acid as a Test Reaction. Eur. J. Inorg. Chem. 2020, 2020, 833–840. [Google Scholar] [CrossRef]
  29. Chaemchuen, S.; Heynderickx, P.M.; Verpoort, F. Kinetic modeling of oleic acid esterification with UiO-66: From intrinsic experimental data to kinetics via elementary reaction steps. Chem. Eng. J. 2020, 394, 124816. [Google Scholar] [CrossRef]
  30. Desidery, L.; Chaemcheun, S.; Yusubov, M.; Verpoort, F. Di-methyl carbonate transesterification with EtOH over MOFs: Basicity and synergic effect of basic and acid active sites. Catal. Commun. 2018, 104, 82–85. [Google Scholar] [CrossRef]
  31. Shaik, M.R.; Adil, S.F.; Alothman, Z.A.; Alduhaish, O.M. Fumarate Based Metal-Organic Framework: An Effective Catalyst for the Transesterification of Used Vegetable Oil. Crystals 2022, 12, 151. [Google Scholar] [CrossRef]
  32. de la Flor, D.; López-Aguado, C.; Paniagua, M.; Morales, G.; Mariscal, R.; Melero, J.A. Defective UiO-66(Zr) as an efficient catalyst for the synthesis of bio jet-fuel precursors via aldol condensation of furfural and MIBK. J. Catal. 2021, 401, 27–39. [Google Scholar] [CrossRef]
  33. Rapeyko, A.; Rodenas, M.; i Xamena, F.X.L. Zr-Containing UiO-66 Metal-Organic Frameworks as Highly Selective Heterogeneous Acid Catalysts for the Direct Ketalization of Levulinic Acid. Adv. Sustain. Syst. 2022, 6, 2100451. [Google Scholar] [CrossRef]
  34. Cirujano, F.; Corma, A.; i Xamena, F.L. Zirconium-containing metal organic frameworks as solid acid catalysts for the esterification of free fatty acids: Synthesis of biodiesel and other compounds of interest. Catal. Today 2015, 257, 213–220. [Google Scholar] [CrossRef]
  35. Cirujano, F.; Corma, A.; i Xamena, F.L. Conversion of levulinic acid into chemicals: Synthesis of biomass derived levulinate esters over Zr-containing MOFs. Chem. Eng. Sci. 2015, 124, 52–60. [Google Scholar] [CrossRef]
  36. Wang, F.; Chen, Z.; Chen, H.; Goetjen, T.A.; Li, P.; Wang, X.; Alayoglu, S.; Ma, K.; Chen, Y.; Wang, T.; et al. Interplay of Lewis and Brønsted Acid Sites in Zr-Based Metal-Organic Frameworks for Efficient Esterification of Biomass-Derived Levulinic Acid. ACS Appl. Mater. Interfaces 2019, 11, 32090–32096. [Google Scholar] [CrossRef]
  37. 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. 2019, 58, 21961–21971. [Google Scholar] [CrossRef]
  38. Abou-Elyazed, A.S.; Sun, Y.; El-Nahas, A.M.; Yousif, A.M. A green approach for enhancing the hydrophobicity of UiO-66(Zr) catalysts for biodiesel production at 298 K. RSC Adv. 2020, 10, 41283–41295. [Google Scholar] [CrossRef]
  39. Desidery, L.; Yusubov, M.S.; Zhuiykov, S.; Verpoort, F. Fully-sulfonated hydrated UiO66 as efficient catalyst for ethyl levulinate production by esterification. Catal. Commun. 2018, 117, 33–37. [Google Scholar] [CrossRef]
  40. Badoei-Dalfard, A.; Shahba, A.; Zaare, F.; Sargazi, G.; Seyedalipour, B.; Karami, Z. Lipase immobilization on a novel class of Zr-MOF/electrospun nanofibrous polymers: Biochemical characterization and efficient biodiesel production. Int. J. Biol. Macromol. 2021, 192, 1292–1303. [Google Scholar] [CrossRef]
  41. Li, H.; Han, Z.; Liu, F.; Li, G.; Guo, M.; Cui, P.; Zhou, S.; Yu, M. Esterification catalyzed by an efficient solid acid synthesized from PTSA and UiO-66(Zr) for biodiesel production. Faraday Discuss. 2021, 231, 342–355. [Google Scholar] [CrossRef] [PubMed]
  42. Xu, Z.; Zhao, G.; Ullah, L.; Wang, M.; Wang, A.; Zhang, Y.; Zhang, S. Acidic ionic liquid based UiO-67 type MOFs: A stable and efficient heterogeneous catalyst for esterification. RSC Adv. 2018, 8, 10009–10016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Peng, W.-L.; Mi, J.; Liu, F.; Xiao, Y.; Chen, W.; Liu, Z.; Yi, X.; Liu, W.; Zheng, A. Accelerating Biodiesel Catalytic Production by Confined Activation of Methanol over High-Concentration Ionic Liquid-Grafted UiO-66 Solid Superacids. ACS Catal. 2020, 10, 11848–11856. [Google Scholar] [CrossRef]
  44. Dai, Q.; Yang, Z.; Li, J.; Cao, Y.; Tang, H.; Wei, X. Zirconium-based MOFs-loaded ionic liquid-catalyzed preparation of biodiesel from Jatropha oil. Renew. Energy 2021, 163, 1588–1594. [Google Scholar] [CrossRef]
  45. Lu, P.; Li, H.; Li, M.; Chen, J.; Ye, C.; Wang, H.; Qiu, T. Ionic liquid grafted NH2-UiO-66 as heterogeneous solid acid catalyst for biodiesel production. Fuel 2022, 324, 124537. [Google Scholar] [CrossRef]
  46. Abou-Elyazed, A.S.; Sun, Y.; El-Nahas, A.; Abdel-Azeim, S.; Sharara, T.; Yousif, A. Solvent-free synthesis and characterization of Ca2+-doped UiO-66(Zr) as heterogeneous catalyst for esterification of oleic acid with methanol: A joint experimental and computational study. Mater. Today Sustain. 2022, 18, 100110. [Google Scholar] [CrossRef]
  47. Zhu, J.; Wang, Z.; Song, X.; Zhao, B.; Li, Y.; Wang, Y. Encapsulating Keggin-H3PW12O40 into UIO-66(Zr) for manufacturing the biodiesel. Micro Nano Lett. 2021, 16, 90–96. [Google Scholar] [CrossRef]
  48. Ma, T.; Liu, D.; Liu, Z.; Xu, J.; Dong, Y.; Chen, G.; Yun, Z. 12-Tungstophosphoric acid-encapsulated metal-organic framework UiO-66: A promising catalyst for the esterification of acetic acid with n-butanol. J. Taiwan Inst. Chem. Eng. 2022, 133, 104277. [Google Scholar] [CrossRef]
  49. Xie, W.; Hu, P.; Yang, X. Cs2.5H0.5PW12O40 Encapsulated in Metal-Organic Framework UiO-66 as Heterogeneous Catalysts for Acidolysis of Soybean Oil. Catal. Lett. 2017, 147, 2772–2782. [Google Scholar] [CrossRef]
  50. 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. 2019, 365, 40–50. [Google Scholar] [CrossRef]
  51. Tan, X.; Zhang, H.; Li, H.; Yang, S. Electrovalent bifunctional acid enables heterogeneously catalytic production of biodiesel by (trans)esterification of non-edible oils. Fuel 2022, 310, 122273. [Google Scholar] [CrossRef]
  52. Li, Y.; Zhang, S.; Li, Z.; Zhang, H.; Li, H.; Yang, S. Green synthesis of heterogeneous polymeric bio-based acid decorated with hydrophobic regulator for efficient catalytic production of biodiesel at low temperatures. Fuel 2022, 329, 125467. [Google Scholar] [CrossRef]
  53. Zhang, Y.; Song, X.; Li, S.; Zhao, B.; Tong, L.; Wang, Y.; Li, Y. Two-step preparation of Keggin-PW12@UIO-66 composite showing high-activity and long-life conversion of soybean oil into biodiesel. RSC Adv. 2021, 11, 38016–38025. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, Q.; Yang, T.; Liu, X.; Yue, C.; Ao, L.; Deng, T.; Zhang, Y. Heteropoly acid-encapsulated metal–organic framework as a stable and highly efficient nanocatalyst for esterification reaction. RSC Adv. 2019, 9, 16357–16365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zhang, Q.; Lei, D.; Luo, Q.; Wang, J.; Deng, T.; Zhang, Y.; Ma, P. Efficient biodiesel production from oleic acid using metal–organic framework encapsulated Zr-doped polyoxometalate nano-hybrids. RSC Adv. 2020, 10, 8766–8772. [Google Scholar] [CrossRef]
  56. Zhang, Q.; Yang, T.; Lei, D.; Wang, J.; Zhang, Y. Efficient Production of Biodiesel from Esterification of Lauric Acid Catalyzed by Ammonium and Silver Co-Doped Phosphotungstic Acid Embedded in a Zirconium Metal-Organic Framework Nanocomposite. ACS Omega 2020, 5, 12760–12767. [Google Scholar] [CrossRef]
  57. Zhang, Q.; Yang, B.; Tian, Y.; Yang, X.; Yu, R.; Wang, J.; Deng, T.; Zhang, Y. Fabrication of silicotungstic acid immobilized on Ce-based MOF and embedded in Zr-based MOF matrix for green fatty acid esterification. Green Process. Synth. 2022, 11, 184–194. [Google Scholar] [CrossRef]
  58. Liu, B.; Shioyama, H.; Akita, T.; Xu, Q. Metal-Organic Framework as a Template for Porous Carbon Synthesis. J. Am. Chem. Soc. 2008, 130, 5390–5391. [Google Scholar] [CrossRef]
  59. Liu, B.; Zhang, X.; Shioyama, H.; Mukai, T.; Sakai, T.; Xu, Q. Converting cobalt oxide subunits in cobalt metal-organic framework into agglomerated Co3O4 nanoparticles as an electrode material for lithium ion battery. J. Power Sources 2010, 195, 857–861. [Google Scholar] [CrossRef]
  60. Lu, N.; Zhang, X.; Yan, X.; Pan, D.; Fan, B.; Li, R. Synthesis of novel mesoporous sulfated zirconia nanosheets derived from Zr-based metal–organic frameworks. CrystEngComm 2020, 22, 44–51. [Google Scholar] [CrossRef]
  61. Li, H.; Wang, Y.; Ma, X.; Guo, M.; Li, Y.; Li, G.; Cui, P.; Zhou, S.; Yu, M. Synthesis of CaO/ZrO2 based catalyst by using UiO–66(Zr) and calcium acetate for biodiesel production. Renew. Energy 2022, 185, 970–977. [Google Scholar] [CrossRef]
  62. Zhang, Q.; Lei, D.; Luo, Q.; Yang, X.; Wu, Y.; Wang, J.; Zhang, Y. MOF-derived zirconia-supported Keggin heteropoly acid nanoporous hybrids as a reusable catalyst for methyl oleate production. RSC Adv. 2021, 11, 8117–8123. [Google Scholar] [CrossRef] [PubMed]
  63. Goda, M.N.; Abdelhamid, H.N.; Said, A.E.-A.A. Zirconium Oxide Sulfate-Carbon (ZrOSO4@C) Derived from Carbonized UiO-66 for Selective Production of Dimethyl Ether. ACS Appl. Mater. Interfaces 2019, 12, 646–653. [Google Scholar] [CrossRef]
  64. Liu, T.; Hong, X.; Liu, G. In Situ Generation of the Cu@3D-ZrOx Framework Catalyst for Selective Methanol Synthesis from CO2/H2. ACS Catal. 2019, 10, 93–102. [Google Scholar] [CrossRef]
  65. Zeng, L.; Wang, Y.; Li, Z.; Song, Y.; Zhang, J.; Wang, J.; He, X.; Wang, C.; Lin, W. Highly Dispersed Ni Catalyst on Metal-Organic Framework-Derived Porous Hydrous Zirconia for CO2 Methanation. ACS Appl. Mater. Interfaces 2020, 12, 17436–17442. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Various heterogeneous catalysts.
Figure 1. Various heterogeneous catalysts.
Bioengineering 09 00700 g001
Figure 2. Schematic diagram of the structure and synthesis of UiO-66.
Figure 2. Schematic diagram of the structure and synthesis of UiO-66.
Bioengineering 09 00700 g002
Figure 3. The suggested structures of UiO-66, UiO-66(COOH)2, and UiO-66(NH2).
Figure 3. The suggested structures of UiO-66, UiO-66(COOH)2, and UiO-66(NH2).
Bioengineering 09 00700 g003
Figure 4. Plausible mechanisms for esterification: (a) UiO-66 catalyst; (b) UiO-66-NH2 acid–base catalyst. (Adapted with permission from Ref. [34]. Copyright 2015, Elsevier.)
Figure 4. Plausible mechanisms for esterification: (a) UiO-66 catalyst; (b) UiO-66-NH2 acid–base catalyst. (Adapted with permission from Ref. [34]. Copyright 2015, Elsevier.)
Bioengineering 09 00700 g004
Figure 5. The suggested structures for AIL and xAIL@NH2-UiO-66 (where x indicates the mass ratio of AIL to NH2-UiO-66).
Figure 5. The suggested structures for AIL and xAIL@NH2-UiO-66 (where x indicates the mass ratio of AIL to NH2-UiO-66).
Bioengineering 09 00700 g005
Figure 6. Reaction mechanism of transesterification on mesoporous sulfated zirconia nanosheets. (Adapted with permission from Ref. [60]. Copyright 2020, Royal Society of Chemistry.)
Figure 6. Reaction mechanism of transesterification on mesoporous sulfated zirconia nanosheets. (Adapted with permission from Ref. [60]. Copyright 2020, Royal Society of Chemistry.)
Bioengineering 09 00700 g006
Figure 7. SEM images for (a) HSiW@UiO-66, (b) HSiW@ZrO2-300, (c) HSiW@ZrO2-400, and (d) HSiW@ZrO2-500. (Adapted with permission from Ref. [62]. Copyright 2021, open access from Royal Society of Chemistry.)
Figure 7. SEM images for (a) HSiW@UiO-66, (b) HSiW@ZrO2-300, (c) HSiW@ZrO2-400, and (d) HSiW@ZrO2-500. (Adapted with permission from Ref. [62]. Copyright 2021, open access from Royal Society of Chemistry.)
Bioengineering 09 00700 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, Q.; Wang, J.; Zhang, S.; Ma, J.; Cheng, J.; Zhang, Y. Zr-Based Metal-Organic Frameworks for Green Biodiesel Synthesis: A Minireview. Bioengineering 2022, 9, 700. https://doi.org/10.3390/bioengineering9110700

AMA Style

Zhang Q, Wang J, Zhang S, Ma J, Cheng J, Zhang Y. Zr-Based Metal-Organic Frameworks for Green Biodiesel Synthesis: A Minireview. Bioengineering. 2022; 9(11):700. https://doi.org/10.3390/bioengineering9110700

Chicago/Turabian Style

Zhang, Qiuyun, Jialu Wang, Shuya Zhang, Juan Ma, Jingsong Cheng, and Yutao Zhang. 2022. "Zr-Based Metal-Organic Frameworks for Green Biodiesel Synthesis: A Minireview" Bioengineering 9, no. 11: 700. https://doi.org/10.3390/bioengineering9110700

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