Next Article in Journal
Investigation of Advanced Oxidation Process in the Presence of TiO2 Semiconductor as Photocatalyst: Property, Principle, Kinetic Analysis, and Photocatalytic Activity
Next Article in Special Issue
Fabrication of Lead-Free Bismuth Based Electroceramic Compositions for High-Energy Storage Density Applications in Electroceramic Capacitors
Previous Article in Journal
Composition and Morphology Modulation of Bimetallic Nitride Nanostructures on Nickel Foams for Efficient Oxygen Evolution Electrocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New 2D Metal-Organic Framework for Photocatalytic Degradation of Organic Dyes in Water

1
Department of Chemistry, School of Science, University of Management and Technology, Lahore 54770, Pakistan
2
School of Chemistry, University of the Punjab, New Campus, Lahore 54590, Pakistan
3
Department of Mechanical and Energy Engineering, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(2), 231; https://doi.org/10.3390/catal13020231
Submission received: 31 December 2022 / Revised: 16 January 2023 / Accepted: 17 January 2023 / Published: 19 January 2023
(This article belongs to the Special Issue Metal-Organic Frameworks in Electro and Photocatalysis)

Abstract

:
Two-dimensional (2D) metal-organic frameworks (MOFs) are fascinating photocatalytic materials because of their unique physical and catalytic properties. Herein, we report a new (E)-4-(3-carboxyacrylamido) benzoic acid [ABA–MA] ligand synthesized under facile conditions. This ABA–MA ligand is further utilized to synthesize a copper-based 2D MOF via the solvothermal process. The resulting 2D MOF is characterized for morphology and electronic structural analysis using advanced techniques, such as proton nuclear magnetic resonance, Fourier-transform infrared spectroscopy, ultraviolet-visible spectroscopy, and scanning electron microscopy. Furthermore, 2D MOF is employed as a photocatalyst for degrading organic dyes, demonstrating the degradation/reduction of methylene blue (MeBl) dye with excellent catalytic/photodegradation activity in the absence of any photosensitizer or cocatalyst. The apparent rate constant (kap) values for the catalytic degradation/reduction of MeBl on the Cu(II)–[ABA-MA] MOF are reported to be 0.0093 min−1, 0.0187 min−1, and 0.2539 min−1 under different conditions of sunlight and NaBH4. The kinetics and stability evaluations reveal the noteworthy photocatalytic potential of the Cu(II)–[ABA–MA] MOF for wastewater treatment. This work offers new insights into the fabrication of new MOFs for highly versatile photocatalytic applications.

1. Introduction

Globally, water contamination has appeared as a severe issue [1]. There are several types of water contaminants, including nitroarenes [2], pesticides [3], heavy metals [4], organic dyes [5], and fertilizers [6]. Among them, dyes are extremely harmful to humans [7], plants [8], and aquatic life [9]. Water contamination is mainly caused by the use of different organic dyes in the leather [10], textile [11], printing [12], pulp [13], and paper [14] industries. The only way to resolve this situation is to ban these industries or eliminate the organic dyes by electro/photodegradation [15]. However, these industries contribute to wider daily life demands and cannot be excluded. Consequently, eliminating organic dyes from water using photocatalysts is the only solution. In order to resolve this issue, a variety of materials, including polymers [16], ligands [17], inorganic nanoparticles [18,19,20], hybrid microgels [21,22,23], and metal-organic frameworks (MOFs) [24,25], have been investigated. The ligands and organic polymers remove these dyes from water through adsorption. However, incomplete elimination of these dyes and a large consumption of adsorbents (ligands and polymers) are the main drawbacks of these adsorbents. Furthermore, it takes a longer time to remove dyes through adsorption. Therefore, ligands and organic polymer adsorbents are ineffective for wastewater purification. The primary disadvantages of using metal nanoparticles are their instability and lack of easily recyclable properties. Therefore, MOFs are promising substances capable of removing toxic organic dyes from water (without any disadvantage) due to their versatile porosity and catalytic properties [26,27,28].
The synthesis and design of MOFs have advanced significantly over the past decades [29]. These materials provide enormous chemical and structural variability as well as a wide range of applications [30,31,32,33]. The essential features affecting the performance of a material are selectivity, permeability, and operating stability [34,35]. Microstructural engineering is also required to control the orientation of MOF crystal formation at the mesoscopic level to improve performance. It is challenging to develop facile synthetic techniques with great accuracy over their microstructural structure at meso- and microscopic levels for optimal performance [35]. The properties of MOFs as an adjustable version (chemical tunability), with unparalleled high surface area, high porosity with flexible pore sizes, robustness, and structural diversity make them suitable for various applications such as sensing, drug delivery, gas storage, magnetic separation, separation, electrochemical energy storage, catalysis, and degradation of toxic organic dyes from aqueous mediums [36,37,38,39,40]. Additionally, MOFs have been utilized as unique precursors for creating inorganic functional materials with previously unimaginable design potential, such as carbons, metal-based compounds, and their composites [41].
Different forms of MOFs, including 2D sheets/bilayers and 3D extended networks, have been designed and used for various applications based on the geometry and connecting properties of the precursors [42,43]. Two-dimensional MOFs have distinctive features and are employed in multiple applications, such as energy conversion and storage, catalysis, and sensing [44]. These materials are excellent candidates for electrochemical applications due to their large specific surface area, tunable structure, porous network, availability of metal active sites, and efficient mass transport properties. Therefore, the appropriate alignment of several donor sites for metal coordination at certain sites is essential to regulate the size, shape, and morphology of MOFs [45,46].
In order to create MOFs, it is necessary to choose the correct metal centers to combine with organic linkers. Copper-based MOFs have received much attention due to the discovery of straight chain secondary building units (SBU), which are advantageous energetically under reaction conditions [47]. Additionally, after activation, the coordinated guest molecules in copper straight chain SBUs can be used in their carboxylic groups to create open sites for Cu(II) ions that are advantageous for applications involving gas storage, adsorption, and catalysis [34,48,49]. Therefore, we selected a unique bidentate ligand to prepare a 2D Cu(II) carboxylate MOF with both carboxyl coordination bonds to copper. A bifunctional carboxylic ligand paired with copper was employed to generate an extended 2D structural framework via copper carboxylate straight chain linkages. In a typical synthesis, a new carboxylic group-containing ligand was synthesized using 4-aminobenzoic acid (4ABA) and maleic anhydride (MA) in acetic acid at room temperature. Then the synthesized ligand (4ABA–MA) was used to form a (Cu(II)–[4ABA–MA]) MOF with copper (II) ions in equimolar condition. The resultant 2D MOF (Cu(II)[4ABA–MA]) was applied as a catalyst in the photocatalytic and catalytic degradation/reduction of methylene blue (MeBl) from an aqueous medium.

2. Results and Discussion

The 4ABA–MA was synthesized by mixing the solutions of 4ABA and MA with gentle stirring in an acidic medium at room temperature (24 °C) (Figure 1) [50]. The acidic medium and low-temperature conditions were used to avoid the cyclization by nitrogen atoms of an amino group. In the acidic condition, the possibility of electron donation from the nitrogen atom of the amino group to maleic anhydride is reduced. Still, the reaction occurs due to the high basicity of the amino group and less stability of maleic anhydride. Upon conversion of the closed ring of maleic anhydride to an open chain after reaction, the electronegativity of the nitrogen atom of the amide group decreases due to the electron withdrawing carbonyl group, which is directly attached to the nitrogen atom in the product (in the form of the amide). Therefore, the nitrogen of the amide group will not attack the carboxylic carbon atom to form a cyclic product in the acidic medium at a low temperature. Then the Cu(II)–[4ABA–MA] MOF was synthesized by the solvothermal process in DMF solvent at a pH of 6.7. Due to the acidic nature of carboxylic groups, deprotonation occurred, and carboxylic groups of 4ABA–MA were converted into carboxylate ions. The electron-donating capacity of 4ABA–MA towards the Cu(II) ions increases due to deprotonation, resulting in MOF formation.
The chemical structural investigations of 4ABA–MA were performed via proton NMR (Figure 2a). The two protons of both carboxylic groups appeared at 11.21 and 10.68 ppm. The carboxylic group, produced by the opening of maleic anhydride, appeared at 11.21 ppm, which is more de-shielded than other carboxylic groups (directly attached to the aromatic ring) due to the resonance effect. Therefore, the protons of both carboxylic groups appeared at different positions. The other protons of aromatic rings and acyclic parts appeared in the 6.96 to 7.96 ppm region. One proton of the amide group merged with the solvent peak in the range of 3.00 to 4.00 ppm. The 1H NMR analysis confirmed the formation of 4ABA–MA. Moreover, the functional group identification and conversion from 4ABA–MA and the Cu(II)–[4ABA–MA] MOF was investigated via FTIR. The O–H, N–H, and C–H stretching peaks of carboxylic groups in 4ABA–MA appeared at 3153 cm−1, 3116 cm−1, and 2867 cm−1, respectively (Figure 2b). The stretching peak of the C=O group appeared at 1722 cm−1 and 1714 cm−1 in both 4ABA–MA and the Cu(II)–[4ABA–MA] MOF. This shifting in the carbonyl stretching peak position of carboxylic groups indicated the involvement of this group in complex formation. The different FTIR spectra of 4ABA–MA and Cu(II)–[4ABA–MA] indicated the formation of the MOF.
The complex formation of 4ABA-MA with copper metal ions was further investigated by UV-visible spectrophotometry (Figure 2c). For this purpose, two separate solutions of both 4ABA–MA and Cu(II)–[4ABA–MA] were prepared in DMF solvent. The λmax of 4ABA–MA appeared at 425 nm and Cu(II)–[4ABA–MA] at 515 nm. This shifting in the value of λmax of 4ABA-MA after adding the copper salt confirmed the formation of a Cu(II)–[4ABA–MA] 2D MOF. The degree of crystallinity of both 4ABA–MA and Cu(II)–[4ABA–MA] was examined with XRD (Figure 2d). The 4ABA-MA do not show any crystallinity, while a high crystallinity is observed in Cu(II)–[4ABA–MA], which indicates the formation of the complex of 4ABA–MA with copper ions as a Cu(II)–[4ABA–MA] MOF.
Furthermore, the morphology of the Cu(II)–[4ABA–MA] MOF was determined by SEM and elemental mapping analysis. The SEM images show the Cu(II)–[4ABA–MA] MOF exists in a 2D hexagonal shape with a smooth surface (Figure 3a,b). Two-dimensional morphology arises from the 2D complexation of 4ABA–MA. There are two carboxylic groups in the 4ABA–MA structure at 180° with respect to each other. Therefore, the 4ABA–MA shows complexation from both sides to form the Cu(II)–[4ABA–MA] MOF, forming a 2D hexagonal structure. Additionally, the elemental mapping analysis of the Cu(II)–[4ABA–MA] MOF was performed, which reveals the composition of the catalysts. Carbon, nitrogen, oxygen, and copper elements are found in the Cu(II)–[4ABA–MA] MOF.
The photocatalytic performance of Cu(II)–[4ABA–MA] was investigated against MeBl via UV-visible spectrophotometry. The catalytic efficiency of Cu(II)–[4ABA–MA] was determined under different conditions, such as in the absence or presence of sunlight and with or without NaBH4 (reductant). Firstly, 0.1 g of Cu(II)–[4ABA–MA] was added to the 100 mL of (0.05 mM) MeBl solution in the absence of NaBH4 and light, and the solution run in a UV-visible spectrophotometer. The UV-visible spectra show a decline in the λmax value for the peak at 656 nm without additional peaks under different conditions (Figure 4a). This spectrum indicates that the Cu(II)–[4ABA–MA] has efficiently degraded the MeBl in 270 min. Similarly, the degradation of MeBl was also monitored in the presence of sunlight (Figure 4b). The peaks of MeBl decreased continuously, resulting in the form of degradation of MeBl. However, it is found that the degradation of MeBl is much faster in the presence of sunlight than in dark conditions. The MeBl is degraded in 120 min in the presence of sunlight compared to 270 min under dark conditions [51]. These results demonstrate the higher photocatalytic efficiency of the Cu(II)–[4ABA–MA].
Additionally, the degradation of MeBl was also examined in sunlight and NaBH4 (Figure 4c) [52,53]. In total, 0.3 g of NaBH4 and 0.1 g of Cu(II)–[4ABA–MA] were added into the (0.05 mM) solution of MeBl and the spectrum recorded. A similar decline is observed in the λmax value at 665 nm, while new peaks appeared (Figure 4c). The decline in the peak at 665 nm and appearance at some other place indicated the reduction of MeBl in the absence of NaBH4. Such peaks did not appear in the absence of NaBH4. These results suggest that the Cu(II)–[4ABA–MA] MOF degraded the MeBl into CO2 and H2O in the absence of NaBH4, but this catalyst reduced the MeBl into some other component. Further research on the separation and purification of these components could provide additional information and mechanisms for this reduction process. The reduction rate of MeBl in the presence of NaBH4 is very high; it takes merely 21 min to completely degrade the dyes in water. Similar behaviour was reported by Melinte et al. [54], where bimetallic (Ag/Au) nanoparticles loaded a crosslinked polymer system for the catalytic and photocatalytic reduction of 4-nitrophenol. Different peaks appeared in the case of catalytic reduction, while there was no peak in the photocatalytic reaction. Therefore, based on similar observations in our results, a similar hypothesis shows the possible mechanism for photocatalytic degradation.
A kinetic study of catalytic and photocatalytic degradation of MeBl was performed to determine the degradation mechanism. The pseudo-first-order kinetic equation was applied to evaluate the apparent rate constant (kap) in the presence and absence of NaBH4 and the presence and absence of sunlight. A plot of ln(At/Ao) vs time determines the reaction rate. The value of ln(At/Ao) decreases slowly with respect to time in the absence of both sunlight and NaBH4 (Figure 5a). This behaviour shows that the Cu(II)–[ABA–MA] MOF has low catalytic efficiency in the absence of sunlight and it increases in the presence of sunlight with a linear relationship between the ln(At/Ao) and time. In the presence of sunlight, the electrons can easily be transferred from ground level to excited results to support the capacity of the electron donor-acceptor process. Therefore, the catalytic performance is higher in the presence of sunlight. The reduction of MeBl rapidly occurred in the presence of NaBH4. It is a fact that NaBH4 is a strong reducing agent. Therefore, the MOF provided the surface area for the catalytic reduction reaction of MeBl, resulting in the reduction of MeBl forming new compounds, which are evidenced by the appearance of new peaks in UV-visible spectra [55,56]. The value of kap was found to be 0.0093 min−1, 0.0187 min−1, and 0.2539 min−1 in the absence of both sunlight and NaBH4 and the presence of only sunlight, and only NaBH4, respectively, for the catalytic degradation/reduction of MeBl in the Cu(II)–[ABA–MA] MOF.
Finally, the structural stability of the Cu(II)–[4ABA–MA] was analyzed by PXRD analysis after photocatalytic evaluation. Similar peaks in PXRD were found before and after catalytic degradation, which demonstrates that the structure of the Cu(II)–[4ABA–MA] remains intact without any changes (Figure 5b). Furthermore, the photocatalytic activity of the Cu(II)–[ABA–MA] MOF was compared with the different photocatalysts on a similar dye degradation system. It is obvious that the photocatalytic activity of Cu(II)–[4ABA–MA] is superior in terms of the apparent rate constant value compared to recently reported catalysts (in Table 1). Additionally, the structural robustness of Cu(II)–[4ABA–MA] is another vital factor for long-term applications. These results indicate that MOF-based photocatalysts are effective for efficient dye degradation due to their versatile structure tunability and porosity and can be utilized for various dyes by altering the ligand or metal centers. Moreover, MOFs show more morphological flexibility and robustness than traditional metal-oxide-based catalysts under different pH conditions. Therefore, these results determine the remarkable potential of the Cu(II)–[4ABA–MA]-based 2D MOF as a dye degradation photocatalyst for wastewater treatment.

3. Materials and Methods

3.1. Materials

Maleic anhydride (MA) (C4H2O3, 99%), 4-aminobenzoic acid (4ABA) (99%), acetic acid (99%), copper(II) sulfate pentahydrate (CuSO4∙5H2O) (98%), and methylene blue (MeBl) were purchased from Sigma-Aldrich. De-ionized water was used for the catalytic and photocatalytic performance of copper(II)–[4ABA–MA].

3.2. Synthesis of 4ABA-MA and Cu(II)–[4ABA–MA]

The equimolar (0.1 mM) solutions of MA and 4ABA were prepared separately in two beakers in 20 mL acetic acid. Then, 4ABA solution was added dropwise into MA solution with gentle stirring at room temperature (24 °C). After mixing the solutions, the reaction was continued for 12 h with gentle stirring to complete the reaction. The yellow color precipitation indicated the formation of the 4ABA–MA ligand. The product was filtered and washed with water to remove the acetic acid molecules. The ligand was dried at 70 °C in an oven. Then this synthesized product was used to form MOF with Cu(II) ions. For this purpose, an equimolar solution of copper sulphate and 4ABA–MA was made in DMF solvent and then mixed at room temperature. This mixture was then added into a Teflon cylindrical autoclave and heated at 110 °C for 24 h. After completing the reaction, the product was cooled at room temperature and dried in an oven. The greenish crystalline Cu(II)–[4ABA–MA] product was formed.

3.3. Characterization Techniques

The synthesized 4ABA was characterized with NMR (400 MHz, Bruker’s Avance Neo Technology, Billerica, MA, USA) technique. The Cary 630 FTIR spectrometer was used to identify the functional groups from MOF. The Bruker D2 Phaser powder X-ray diffraction (PXRD) technique was used for crystal structure analysis of 2D MOF. The UV-Vis spectrophotometer (V–780 UV-Vis/NIR) was used at room temperature. The MOF was also analyzed by Field Emission Scanning Electron Microscope (SEM) (FEI Nova 450 NanoSEM).

3.4. Electrocatalytic Application

The synthesized Cu(II)–[4ABA–MA] was used as a catalyst for the degradation of MeBl dyes from an aqueous medium. The catalytic efficiency of Cu(II)–[4ABA–MA] MOF was observed with and without NaBH4. The performance of MOF was also investigated in the presence and absence of light in an aqueous medium. The same amount (0.01 g) of MOF was used for the catalytic degradation of 100 mL (0.05 mM) MeBl solution.

4. Conclusions

In this study, we report a new ligand synthesized using ABA and MA in the acid medium under ambient conditions. The 2D Cu(II)–[4ABA–MA] MOF was prepared using the newly synthesized ligand by the solvothermal process and employed as a photocatalyst in wastewater treatment. The 2D MOF showed excellent performance as a dye degradation catalyst in aqueous media, demonstrating the highly effective photocatalyst. Kinetic studies show that the Cu(II)–[ABA–MA] MOF has promising photocatalytic properties for the degradation/reduction of MeBl in the Cu(II)–[ABA–MA] MOF. Structural stability evaluation further demonstrates the practical applications of the Cu(II)–[ABA–MA] MOF in wastewater treatment. This study presents insightful designs and applications of Cu-based MOFs for waste treatment under mild conditions.

Author Contributions

Conceptualization and writing, M.A., S.Z. and U.F.; methodology, A.R.; software, validation, Z.H.F. and M.F.; formal analysis, M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data will be provided on request.

Acknowledgments

Muhammad Arif is thankful to the University of Management and Technology, Lahore, Pakistan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwarzenbach, R.P.; Egli, T.; Hofstetter, T.B.; Von Gunten, U.V.; Wehrli, B. Global Water Pollution and Human Health. Annu. Rev. Environ. Resour. 2010, 35, 109–136. [Google Scholar] [CrossRef]
  2. Ma, F.; Yuan, G.; Meng, L.; Oda, Y.; Hu, J. Contributions of flumequine and nitroarenes to the genotoxicity of river and ground waters. Chemosphere 2012, 88, 476–483. [Google Scholar] [CrossRef]
  3. Agrawal, A.; Pandey, R.S.; Sharma, B. Water Pollution with Special Reference to Pesticide Contamination in India. J. Water Resour. Prot. 2010, 2, 432–448. [Google Scholar] [CrossRef] [Green Version]
  4. Zamora-Ledezma, C.; Negrete-Bolagay, D.; Figueroa, F.; Zamora-Ledezma, E.; Ni, M.; Alexis, F.; Guerrero, V.H. Heavy metal water pollution: A fresh look about hazards, novel and conventional remediation methods. Environ. Technol. Innov. 2021, 22, 101504. [Google Scholar] [CrossRef]
  5. Ismail, M.; Akhtar, K.; Khan, M.I.; Kamal, T.; Khan, M.A.; Asiri, A.M.; Seo, J.; Khan, S.B. Pollution, Toxicity and Carcinogenicity of Organic Dyes and their Catalytic Bio-Remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef]
  6. Savci, S. An Agricultural Pollutant: Chemical Fertilizer. Int. J. Environ. Sci. Dev. 2012, 3, 73–80. [Google Scholar] [CrossRef] [Green Version]
  7. Chung, K.-T. Azo dyes and human health: A review. J. Environ. Sci. Health Part C 2016, 34, 233–261. [Google Scholar] [CrossRef] [PubMed]
  8. Chandanshive, V.V.; Kadam, S.K.; Khandare, R.V.; Kurade, M.B.; Jeon, B.-H.; Jadhav, J.P.; Govindwar, S.P. In situ phytoremediation of dyes from textile wastewater using garden ornamental plants, effect on soil quality and plant growth. Chemosphere 2018, 210, 968–976. [Google Scholar] [CrossRef]
  9. Hassaan, M.; El Nemr, A. Health and Environmental Impacts of Dyes: Mini Review. Am. J. Environ. Sci. Eng. 2017, 1, 64. [Google Scholar] [CrossRef]
  10. Khayet, M.; Zahrim, A.; Hilal, N. Modelling and optimization of coagulation of highly concentrated industrial grade leather dye by response surface methodology. Chem. Eng. J. 2011, 167, 77–83. [Google Scholar] [CrossRef]
  11. Benkhaya, B.; Harfi, S.E.; Harfi, A.E. Classifications, Properties and Applications of Textile Dyes: A Review. Appl. J. Environ. Eng. Sci. 2017, 3, 311–320. [Google Scholar] [CrossRef]
  12. Gastaldi, M.; Cardano, F.; Zanetti, M.; Viscardi, G.; Barolo, C.; Bordiga, S.; Magdassi, S.; Fin, A.; Roppolo, I. Functional Dyes in Polymeric 3D Printing: Applications and Perspectives. ACS Mater. Lett. 2021, 3, 1–17. [Google Scholar] [CrossRef]
  13. Hermosilla, D.; Merayo, N.; Gascó, A.; Blanco, Á. The application of advanced oxidation technologies to the treatment of effluents from the pulp and paper industry: A review. Environ. Sci. Pollut. Res. 2015, 22, 168–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liao, J.; Zaman, S.; Wang, Y.; Yang, M.; Yang, L.; Chen, M.; Wang, H. Improved Reversal Tolerant Properties of IrO(X) as an Anode Electrocatalyst in Proton Exchange Membrane Fuel Cells. ACS Appl. Mater. Interfaces 2023, 11, 1–9. [Google Scholar] [CrossRef]
  15. Haseeb, M.; Haouas, I.; Nasih, M.; Mihardjo, L.W.; Jermsittiparsert, K. Asymmetric impact of textile and clothing manufacturing on carbon-dioxide emissions: Evidence from top Asian economies. Energy 2020, 196, 117094. [Google Scholar] [CrossRef]
  16. Blanco, S.P.D.M.; Scheufele, F.B.; Módenes, A.N.; Espinoza-Quiñones, F.R.; Marin, P.; Kroumov, A.D.; Borba, C.E. Kinetic, equilibrium and thermodynamic phenomenological modeling of reactive dye adsorption onto polymeric adsorbent. Chem. Eng. J. 2017, 307, 466–475. [Google Scholar] [CrossRef]
  17. Tsai, M.-J.; Wu, J.-Y. Synthesis, Structure, and Dye Adsorption Properties of a Nickel(II) Coordination Layer Built from d-Camphorate and Bispyridyl Ligands. Polymers 2017, 9, 661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Rahman, Q.I.; Ahmad, M.; Misra, S.K.; Lohani, M. Effective photocatalytic degradation of rhodamine B dye by ZnO nanoparticles. Mater. Lett. 2013, 91, 170–174. [Google Scholar] [CrossRef]
  19. Zaman, S.; Huang, L.; Douka, A.I.; Yang, H.; You, B.; Xia, B.Y. Oxygen Reduction Electrocatalysts toward Practical Fuel Cells: Progress and Perspectives. Angew. Chem. Int. Ed. 2021, 60, 17832–17852. [Google Scholar] [CrossRef]
  20. Arif, M.; Shahid, M.; Irfan, A.; Nisar, J.; Wang, X.; Batool, N.; Ali, M.; Farooqi, Z.H.; Begum, R.; Farooqi, R. Begum, Extraction of copper ions from aqueous medium by microgel particles for in-situ fabrication of copper nanoparticles to degrade toxic dyes, Zeitschrift Für Phys. Chemie 2022, 236, 1219–1241. [Google Scholar] [CrossRef]
  21. Huang, L.; Zaman, S.; Tian, X.; Wang, Z.; Fang, W.; Xia, B.Y. Advanced Platinum-Based Oxygen Reduction Electrocatalysts for Fuel Cells. Accounts Chem. Res. 2021, 54, 311–322. [Google Scholar] [CrossRef]
  22. Arif, M.; Shahid, M.; Irfan, A.; Nisar, J.; Wu, W.; Farooqi, Z.H.; Begum, R. Polymer microgels for the stabilization of gold nanoparticles and their application in the catalytic reduction of nitroarenes in aqueous media. RSC Adv. 2022, 12, 5105–5117. [Google Scholar] [CrossRef]
  23. Arif, M. Complete life of cobalt nanoparticles loaded into cross-linked organic polymers: A review. RSC Adv. 2022, 12, 15447–15460. [Google Scholar] [CrossRef]
  24. Zhang, M.; Wang, L.; Zeng, T.; Shang, Q.; Zhou, H.; Pan, Z.; Cheng, Q. Two pure MOF-photocatalysts readily prepared for the degradation of methylene blue dye under visible light. Dalton Trans. 2018, 47, 4251–4258. [Google Scholar] [CrossRef] [PubMed]
  25. Zhu, W.; Yang, X.-Y.; Li, Y.-H.; Li, J.-P.; Wu, D.; Gao, Y.; Yi, F.-Y. A novel porous molybdophosphate-based FeII,III-MOF showing selective dye degradation as a recyclable photocatalyst. Inorg. Chem. Commun. 2014, 49, 159–162. [Google Scholar] [CrossRef]
  26. Chen, S.; Zhang, Z.; Jiang, W.; Zhang, S.; Zhu, J.; Wang, L.; Ou, H.; Zaman, S.; Tan, L.; Zhu, P.; et al. Engineering Water Molecules Activation Center on Multisite Electrocatalysts for Enhanced CO2 Methanation. J. Am. Chem. Soc. 2022, 144, 12807–12815. [Google Scholar] [CrossRef] [PubMed]
  27. Qin, L.; Liang, F.; Li, Y.; Wu, J.; Guan, S.; Wu, M.; Xie, S.; Luo, M.; Ma, D. A 2D Porous Zinc-Organic Framework Platform for Loading of 5-Fluorouracil. Inorganics 2022, 10, 202. [Google Scholar] [CrossRef]
  28. Dong, X.; Shi, Z.; Li, D.; Li, Y.; An, N.; Shang, Y.; Sakiyama, H.; Muddassir, M.; Si, C. The regulation research of topology and magnetic exchange models of CPs through Co(II) concentration adjustment. J. Solid State Chem. 2023, 318, 123713. [Google Scholar] [CrossRef]
  29. Radwan, A.; Jin, H.; He, D.; Mu, S. Design Engineering, Synthesis Protocols, and Energy Applications of MOF-Derived Electrocatalysts. Nano-Micro Lett. 2021, 13, 1–32. [Google Scholar] [CrossRef]
  30. Hao, M.; Qiu, M.; Yang, H.; Hu, B.; Wang, X. Recent advances on preparation and environmental applications of MOF-derived carbons in catalysis. Sci. Total. Environ. 2020, 760, 143333. [Google Scholar] [CrossRef]
  31. Zaman, S.; Su, Y.; Dong, C.; Qi, R.; Huang, L.; Qin, Y.; Huang, Y.; Li, F.; You, B.; Guo, W.; et al. Scalable Molten Salt Synthesis of Platinum Alloys Planted in Metal-Nitrogen-Graphene for Efficient Oxygen Reduction. Angew. Chem. 2021, 134, e202115835. [Google Scholar] [CrossRef]
  32. Zhang, M.; Bosch, M.; Iii, T.G.; Zhou, H.-C. Rational design of metal-organic frameworks with anticipated porosities and functionalities. Crystengcomm 2014, 16, 4069–4083. [Google Scholar] [CrossRef] [Green Version]
  33. Xie, W.; Cui, D.; Zhang, S.-R.; Xu, Y.-H.; Jiang, D.-L. Iodine capture in porous organic polymers and metal-organic frameworks materials. Mater. Horizons 2019, 6, 1571–1595. [Google Scholar] [CrossRef]
  34. Yang, S.; Karve, V.V.; Justin, A.; Kochetygov, I.; Espín, J.; Asgari, M.; Trukhina, O.; Sun, D.T.; Peng, L.; Queen, W.L. Enhancing MOF performance through the introduction of polymer guests. Coord. Chem. Rev. 2020, 427, 213525. [Google Scholar] [CrossRef]
  35. Niu, H.; Xia, C.; Huang, L.; Zaman, S.; Maiyalagan, T.; Guo, W.; You, B.; Xia, B.Y. Rational design and synthesis of one-dimensional platinum-based nanostructures for oxygen-reduction electrocatalysis. Chin. J. Catal. 2022, 43, 1459–1472. [Google Scholar] [CrossRef]
  36. Kirchon, A.; Feng, L.; Drake, H.F.; Joseph, E.A.; Zhou, H.-C. From fundamentals to applications: A toolbox for robust and multifunctional MOF materials. Chem. Soc. Rev. 2018, 47, 8611–8638. [Google Scholar] [CrossRef]
  37. Zaman, S.; Tian, X.; Su, Y.-Q.; Cai, W.; Yan, Y.; Qi, R.; Douka, A.I.; Chen, S.; You, B.; Liu, H.; et al. Direct integration of ultralow-platinum alloy into nanocarbon architectures for efficient oxygen reduction in fuel cells. Sci. Bull. 2021, 66, 2207–2216. [Google Scholar] [CrossRef]
  38. Ali, H.; Zaman, S.; Majeed, I.; Kanodarwala, F.K.; Nadeem, M.A.; Stride, J.A. Porous Carbon/rGO Composite: An Ideal Support Material of Highly Efficient Palladium Electrocatalysts for the Formic Acid Oxidation Reaction. Chemelectrochem 2017, 4, 3126–3133. [Google Scholar] [CrossRef]
  39. Liu, W.; Yan, Q.; Xia, C.; Wang, X.; Kumar, A.; Wang, Y.; Liu, Y.; Pan, Y.; Liu, J. Recent advances in cell membrane coated metal-organic frameworks (MOFs) for tumor therapy. J. Mater. Chem. B 2021, 9, 4459–4474. [Google Scholar] [CrossRef]
  40. Li, M.; Yin, S.; Lin, M.; Chen, X.; Pan, Y.; Peng, Y.; Sun, J.; Kumar, A.; Liu, J. Current status and prospects of metal-organic frameworks for bone therapy and bone repair. J. Mater. Chem. B 2022, 10, 5105–5128. [Google Scholar] [CrossRef]
  41. Yu, J.; Mu, C.; Yan, B.; Qin, X.; Shen, C.; Xue, H.; Pang, H. Nanoparticle/MOF composites: Preparations and applications. Mater. Horizons 2017, 4, 557–569. [Google Scholar] [CrossRef]
  42. Ahmad, N.; Younus, H.A.; Chughtai, A.H.; Van Hecke, K.; Khattak, Z.A.K.; Gaoke, Z.; Danish, M.; Verpoort, F. Synthesis of 2D MOF having potential for efficient dye adsorption and catalytic applications. Catal. Sci. Technol. 2018, 8, 4010–4017. [Google Scholar] [CrossRef]
  43. Zhao, S. A novel 3D MOF with rich lewis basic sites as a base catalysis toward knoevenagel condensation reaction. J. Mol. Struct. 2018, 1167, 11–15. [Google Scholar] [CrossRef]
  44. Zhu, H.; Liu, D. The synthetic strategies of metal-organic framework membranes, films and 2D MOFs and their applications in devices. J. Mater. Chem. A 2019, 7, 21004–21035. [Google Scholar] [CrossRef]
  45. Xue, Y.; Zhao, G.; Yang, R.; Chu, F.; Chen, J.; Wang, L.; Huang, X. 2D metal-organic framework-based materials for electrocatalytic, photocatalytic and thermocatalytic applications. Nanoscale 2021, 13, 3911–3936. [Google Scholar] [CrossRef]
  46. Wang, H.; Wang, W.; Zaman, S.; Yu, Y.; Wu, Z.; Liu, H.; Xia, B.Y. Dicyandiamide and iron-tannin framework derived nitrogen-doped carbon nanosheets with encapsulated iron carbide nanoparticles as advanced pH-universal oxygen reduction catalysts. J. Colloid Interface Sci. 2018, 530, 196–201. [Google Scholar] [CrossRef]
  47. Raoof, J.-B.; Hosseini, S.R.; Ojani, R.; Mandegarzad, S. MOF-derived Cu/nanoporous carbon composite and its application for electro-catalysis of hydrogen evolution reaction. Energy 2015, 90, 1075–1081. [Google Scholar] [CrossRef]
  48. Ding, M.; Cai, X.; Jiang, H.-L. Improving MOF stability: Approaches and applications. Chem. Sci. 2019, 10, 10209–10230. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, M.; Zi, Y.; Zhu, J.; Huang, W.; Zhang, Z.; Zhang, H. Construction of super-hydrophobic PDMS@MOF@Cu mesh for reduced drag, anti-fouling and self-cleaning towards marine vehicle applications. Chem. Eng. J. 2021, 417, 129265. [Google Scholar] [CrossRef]
  50. Zaman, S.; Wang, M.; Liu, H.; Sun, F.; Yu, Y.; Shui, J.; Chen, M.; Wang, H. Carbon-based catalyst supports for oxygen reduction in proton-exchange membrane fuel cells. Trends Chem. 2022, 4, 886–906. [Google Scholar] [CrossRef]
  51. Samuel, M.S.; Savunthari, K.V.; Ethiraj, S. Synthesis of a copper (II) metal-organic framework for photocatalytic degradation of rhodamine B dye in water. Environ. Sci. Pollut. Res. 2021, 28, 40835–40843. [Google Scholar] [CrossRef]
  52. Arif, M.; Shahid, M.; Irfan, A.; Wang, X.; Noor, H.; Farooqi, Z.H.; Begum, R. Catalytic degradation of organic dyes using Au-poly(styrene@N-isopropylmethacrylamide) hybrid microgels. Inorg. Chem. Commun. 2022, 144, 109870. [Google Scholar] [CrossRef]
  53. Zou, J.; Huang, H.; Zaman, S.; Yao, K.; Xing, S.; Chen, M.; Wang, H.; Wang, M. Enhanced electrochemical hydrogen compression performance with a gradient water-retaining hybrid membrane. Chem. Eng. J. 2023, 457, 141113–141123. [Google Scholar] [CrossRef]
  54. Melinte, V.; Stroea, L.; Buruiana, T.; Chibac, A.L. Photocrosslinked hybrid composites with Ag, Au or Au-Ag NPs as visible-light triggered photocatalysts for degradation/reduction of aromatic nitroderivatives. Eur. Polym. J. 2019, 121, 109289. [Google Scholar] [CrossRef]
  55. Javed, M.; Abid, M.; Hussain, S.; Shahwar, D.; Arshad, S.; Ahmad, N.; Arif, M.; Khan, H.; Nadeem, S.; Raza, H.; et al. Synthesis, characterization and photocatalytic applications of s-doped graphitic carbon nitride nanocomposites with nickel doped zinc oxide nanoparticles. Dig. J. Nanomater. Biostruct. 2020, 15, 1097–1105. [Google Scholar]
  56. Arif, M. Extraction of iron (III) ions by core-shell microgel for in situ formation of iron nanoparticles to reduce harmful pollutants from water. J. Environ. Chem. Eng. 2023, 11, 109270. [Google Scholar] [CrossRef]
  57. Yang, C.; Dong, W.; Cui, G.; Zhao, Y.; Shi, X.; Xia, X.; Tang, B.; Wang, W. Highly efficient photocatalytic degradation of methylene blue by P2ABSA-modified TiO2 nanocomposite due to the photosensitization synergetic effect of TiO2 and P2ABSA. RSC Adv. 2017, 7, 23699–23708. [Google Scholar] [CrossRef] [Green Version]
  58. Vallejo, W.; Cantillo, A.; Díaz-Uribe, C. Methylene Blue Photodegradation under Visible Irradiation on Ag-Doped ZnO Thin Films. Int. J. Photoenergy 2020, 2020, 1627498. [Google Scholar] [CrossRef] [Green Version]
  59. Balcha, A.; Yadav, O.P.; Dey, T. Photocatalytic degradation of methylene blue dye by zinc oxide nanoparticles obtained from precipitation and sol-gel methods. Environ. Sci. Pollut. Res. 2016, 23, 25485–25493. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of Cu(II) synthesis–[ABA–MA] MOF.
Figure 1. Schematic illustration of Cu(II) synthesis–[ABA–MA] MOF.
Catalysts 13 00231 g001
Figure 2. Chemical structural characterization of 2D MOF: (a) 1H NMR spectrum of 4ABA–MA in DMSO and (b) FTIR of 4ABA–MA and Cu(II)–[4ABA–MA] MOF. (c) UV-visible spectrum of 4ABA–MA and Cu(II)–[4ABA–MA] (d) PXRDof 4ABA–MA and Cu(II)–[4ABA–MA] MOF.
Figure 2. Chemical structural characterization of 2D MOF: (a) 1H NMR spectrum of 4ABA–MA in DMSO and (b) FTIR of 4ABA–MA and Cu(II)–[4ABA–MA] MOF. (c) UV-visible spectrum of 4ABA–MA and Cu(II)–[4ABA–MA] (d) PXRDof 4ABA–MA and Cu(II)–[4ABA–MA] MOF.
Catalysts 13 00231 g002
Figure 3. Morphology analysis. (a,b) SEM images and (c) elemental mapping analysis of Cu(II)–[4ABA–MA].
Figure 3. Morphology analysis. (a,b) SEM images and (c) elemental mapping analysis of Cu(II)–[4ABA–MA].
Catalysts 13 00231 g003
Figure 4. Catalytic and photocatalytic degradation/reduction of MeBl on Cu(II)–[ABA–MA] MOF; (a) in the absence of both sunlight and NaBH4, (b) in the presence of sunlight but the absence of NaBH4, (c) in the presence of NaBH4 but the absence of sunlight.
Figure 4. Catalytic and photocatalytic degradation/reduction of MeBl on Cu(II)–[ABA–MA] MOF; (a) in the absence of both sunlight and NaBH4, (b) in the presence of sunlight but the absence of NaBH4, (c) in the presence of NaBH4 but the absence of sunlight.
Catalysts 13 00231 g004
Figure 5. Photocatalytic performance of Cu(II)–[ABA–MA] MOF (a) Degradation kinetics under different conditions (the black line represents the absence of both sunlight and NaBH4 and the red line represents the presence of sunlight but the absence of NaBH4, while the blue line represents the presence of NaBH4 but the absence of sunlight). (b) XRD spectra of Cu(II)–[ABA–MA] MOF before (red color) and after (green color) photocatalytic degradation of MeBl.
Figure 5. Photocatalytic performance of Cu(II)–[ABA–MA] MOF (a) Degradation kinetics under different conditions (the black line represents the absence of both sunlight and NaBH4 and the red line represents the presence of sunlight but the absence of NaBH4, while the blue line represents the presence of NaBH4 but the absence of sunlight). (b) XRD spectra of Cu(II)–[ABA–MA] MOF before (red color) and after (green color) photocatalytic degradation of MeBl.
Catalysts 13 00231 g005
Table 1. Comparison of photocatalytic activity of Cu(II)–[ABA–MA] MOF with other systems.
Table 1. Comparison of photocatalytic activity of Cu(II)–[ABA–MA] MOF with other systems.
CatalystAbbreviationsSubstrateApparent Rate Constant (kap) (min−1)Refs
TiO2 particlesTiO2MeBl0.0021[57]
TiO2 composite TiO2–P(ABSA)MeBl0.0138[57]
Ag–doped ZnOAg@ZnOMeBl0.0059[58]
Zinc oxideZnOMeBl0.0084[59]
Copper (II)–[4-benzoic acid-maleic anhydride]Cu(II)–[4ABA–MA]MeBl0.0187 This 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

Arif, M.; Fatima, U.; Rauf, A.; Farooqi, Z.H.; Javed, M.; Faizan, M.; Zaman, S. A New 2D Metal-Organic Framework for Photocatalytic Degradation of Organic Dyes in Water. Catalysts 2023, 13, 231. https://doi.org/10.3390/catal13020231

AMA Style

Arif M, Fatima U, Rauf A, Farooqi ZH, Javed M, Faizan M, Zaman S. A New 2D Metal-Organic Framework for Photocatalytic Degradation of Organic Dyes in Water. Catalysts. 2023; 13(2):231. https://doi.org/10.3390/catal13020231

Chicago/Turabian Style

Arif, Muhammad, Urooj Fatima, Abdul Rauf, Zahoor Hussain Farooqi, Mohsin Javed, Muhammad Faizan, and Shahid Zaman. 2023. "A New 2D Metal-Organic Framework for Photocatalytic Degradation of Organic Dyes in Water" Catalysts 13, no. 2: 231. https://doi.org/10.3390/catal13020231

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