Next Article in Journal
Optimization of the Composition of Mesoporous Polymer–Ceramic Nanocomposite Granules for Bone Regeneration
Previous Article in Journal
Sequential Esterification—Diels-Alder Reactions for Improving Pine Rosin Durability within Road Marking Paint
Previous Article in Special Issue
Molecular Design of Porous Organic Polymer-Derived Carbonaceous Electrocatalysts for Pinpointing Active Sites in Oxygen Reduction Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Study on Gas Sorption and Iodine Uptake of a Metal-Organic Framework Based on Curcumin

1
School of Environmental Science and Engineering, Yancheng Institute of Technology, Yancheng 224051, China
2
State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130021, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(13), 5237; https://doi.org/10.3390/molecules28135237
Submission received: 31 May 2023 / Revised: 24 June 2023 / Accepted: 1 July 2023 / Published: 6 July 2023
(This article belongs to the Special Issue Potential Applications of Functional Porous Organic Frameworks)

Abstract

:
Medi-MOF-1 is a highly porous Metal-Organic framework (MOF) constructed from Zn(II) and curcumin. The obtained crystal was characterized using powder X-ray diffraction (PXRD) and scanning electron microscopy (SEM). A micrometer-sized crystal with similar morphology was successfully obtained using the solvothermal method. Thanks to its high surface area, good stability, and abound pores, the as-synthesized medi-MOF-1 could be used as a functional porous material to adsorb different gases (H2, CO2, CH4, and N2) and iodine (I2). The activated sample exhibited a high I2 adsorption ability of 1.936 g g–1 at room temperature via vapor diffusion. Meanwhile, the adsorbed I2 could be released slowly in ethanol, confirming the potential application for I2 adsorption.

1. Introduction

With the rapid growth of global energy demand, the pursuit of energy storage technology has drawn special attention from scientists in recent years [1,2]. Compared to coal and petroleum, gaseous fuels are more friendly to the environment. However, the transportation and storage of gaseous fuels is a major challenge for researchers. Therefore, functional porous materials to capture or separate gas have been investigated, such as zeolite [3,4,5], Metal-Organic frameworks [6,7,8], and porous organic polymers [9,10]. Among these materials, Metal-Organic frameworks (MOFs), constructed from secondary building units (SBUs) and organic linkers, have been widely researched in the past decades. Due to their high porosities, tunable pore sizes, etc., the unique structural advantage of MOFs has been widely pursued for gas storage and separation [11,12,13,14], catalysis [15], chemical sensing [16,17,18], and so on. Hydrogen is considered as one of the alternative energy sources for fossil fuels. With a large surface area and tunable pore structure, MOFs can also provide unsaturated metal sites for hydrogen sorption [19,20,21]. At present, growing efforts to develop MOFs that efficiently capture or separate CO2 is desired [22,23,24,25]. Compared to traditional methods, MOFs based on adsorption to capture or separate CO2 can reduce energy consumption, showing great advantages in these technologies. Thus, constructing an economical and preferable tunable pore structure and pores is highly desirable.
Moreover, MOFs also exhibit potential applications for toxic waste elimination, such as heavy metal [26], uranium [27], iodine [28,29,30,31], and so on. Radioactive iodine 129I and 131I have a long half-life (1.57 × 107 years), which compounds damage the environment and human beings. How to dispose the nuclear waste timely and effectively has become an important issue that needs to be addressed. Although zerovalent iron [32], zeolite [33], and functionalized clays [34] have been used for radioiodine capture, their low absorption capacity and less interactive sites with iodine have limited the applications. Thus, MOFs with designable architecture and excellent properties have been synthesized for iodine capture. There are two kinds of MOFs for iodine adsorption, including non-iodine MOFs and iodine-templated MOFs [35,36,37]. The presence of iodide groups in the framework can affect the iodine uptake [38,39]. For example, [(ZnI2)3(tpt)2] was constructed with ZnI2 nodes with tpt (2,4,6-tris(4-pyridyl)-1,3,5-triazine), which has a good I2 loading amount of 173 wt.% at room temperature, which is similar to reported for Cu-BTC (175 wt.%) [40]. In addition, non-iodine MOFs have been developed and utilized in the field to capture I2. [Zr6O4(OH)4(edb)6]n can uptake iodine by means of chemi- and physisorption [41]. Although the presence of the iodide group in the framework can affect the iodine uptake, the structure of the organic ligands is too complex, which can make it difficult to obtain in the synthetic reaction. Hence, choosing a simple ligand and reaction process to construct porous MOFs for gas and I2 sorption is highly needed.
In our previous work [42], medi-MOF-1 was successfully synthesized using structurally symmetric ligand curcumin with the Zn(II) ion. In this study, we report on micrometer-sized porous medi-MOF-1 via large-scale reactions (Scheme 1). Subsequently, we studied the adsorption properties of medi-MOF-1 for H2, CO2, CH4, and N2. It is revealed that medi-MOF-1 can adsorb 1.57 wt.% H2 at 77 K and 1 bar, and displayed commendable CO2 adsorption and selectivity for CO2 over CH4 and N2 at 273 K. Furthermore, it is worth noting that medi-MOF-1 exhibits an outstanding I2 adsorption capacity of 1.936 g g−1. It is also revealed that the I2 sorption process of medi-MOF-1 is reversible.

2. Results

2.1. Physicochemical Properties of Medi-MOF-1

As reported previously by us, medi-MOF-1 crystallizes in the trigonal chiral space group P3221. It is constructed by trinuclear clusters and curcumin ligands, leading to a three-dimensional (3D) porous coordination framework (Scheme 1). In this work, we synthesized medi-MOF-1 by expanding the reaction by 10 times. Powder X-ray diffraction (PXRD) of medi-MOF-1 confirmed the phase purity of the bulk crystalline materials due to the same PXRD pattern with the simulated data (Figure 1b). The pore structure properties of medi-MOF-1 were characterized at 77 K. The N2 adsorption–desorption isotherm of type-I adsorption curves with a capillary in the low P/P0 region. It confirms that the as-synthesized medi-MOF-1 is microporous. The adsorption isotherm data were fitted to the Langmuir equation and gave a surface area of 563 m2 g−1 (BET surface area: 475 m2 g−1). This surface area is similar to nanosized medi-MOF-1 particles but far below the value of 2675 m2 g−1 reported large size of the crystal [42]. The decreased size and crystallinity may be the main reason for reducing the BET of medi-MOF-1 [43]. Subsequently, scanning electron microscopy (SEM) technologies have been widely used to study the morphology of nanoparticles. Herein, SEM images were recorded for inspecting the morphology and structure of as-synthesized medi-MOF-1. As seen in Figure 1c,d, SEM images show that the prepared solid samples are agglomerated with small nanocrystals which have good uniformity and dimensional consistency. SEM images showed rod-like crystals of medi-MOF-1 with a diameter of 0.2 μm and length of 1 μm which had a similar morphology to that of larger crystals [42]. According to these results, the as-synthesized medi-MOF-1 can be successfully synthesized as crystalline powder materials.

2.2. Gas Sorption Properties

Much effort has been devoted to hydrogen storage since hydrogen is considered to be an excellent alternative energy source. In order to test the hydrogen uptake of medi-MOF-1, the as-synthesized medi-MOF-1 was activated after soaking in the solvent of CH2Cl2 for 2 days [44] and heating at 100 °C under vacuum for 8 h. The H2 sorption experiments of activated samples were measured, which showed that the H2 uptake of medi-MOF-1 is as high as 1.57 wt.% at 77 K and 1 atm (Figure 2a), which is similar to bio-MOF-11 constructed from biomolecules solely [45]. We calculated the isosteric heats of adsorption (Qst) of H2 using the Clausius equation following the fitting of the isotherm data at 77 and 87 K using a virial equation. The initial Qst value for H2 of medi-MOF-1 is calculated to be −6.25 kJ mol−1 at zero coverage (Figure 2b). Furthermore, the initial Qst value for H2 of medi-MOF-1 is smaller than that of bio-MOF-11 (−13 kJ mol−1) [46,47]. These results may be attributed to the more metal clusters of bio-MOF-11, for the force of metal clusters on hydrogen is greater than that of the benzene ring on hydrogen. Additionally, the M2(olz) materials are bioactive frameworks with similar frameworks exhibiting potential H2 storage capacities [48]. The relevant BET data and H2 adsorption capacities of the selected MOFs constructed with biomolecules or drug molecules are summarized and tabulated in Table S1. The isosteric heat of Zn2(olz) adsorption is similar to medi-MOF-1, which is lower than other M2(olz) frameworks in the series due to fewer open metal sites in the activated materials.
As we know, the separation of CO2 to CH4 and N2 by porous materials is favorable in the environment. In this text, we measured the adsorption ability of different gases in medi-MOF-1. At 273 K and 1 bar, medi-MOF-1 exhibits higher uptake of CO2 than CH4 and N2—which is 34.7 cm3 g−1 (1.55 mmol g−1). The datapoint is lower than those of bio-MOFs constructed with biomolecules [49]. The reason for this phenomenon may be the fewer adsorption sites in the framework of medi-MOF-1. The relevant BET data and CO2 adsorption capacities of the selected bio-MOFs are summarized and tabulated in Table 1. The maximum uptakes of CH4 and N2 are 13.9 cm3 g−1 and 5.95 cm3 g−1 at 273 K and 1 bar, respectively (Figure 3a), which are lower than that of CO2 in the same conditions. Especially, the uptake of CO2 for medi-MOF-1 is 2.5 times as large as CH4 and 5.8 times as large as that of N2. The adsorption selectivity of CO2 relative to CH4 and N2 was calculated using Henry’s law before 0.1 bar. Based on Henry’s law, the material shows CO2 over N2 or CH4 adsorption selectivity (Figure 3b), ranking medi-MOF-1 as better porous adsorbents constructed from biomolecules for separating CO2 from N2 [50,51,52].

2.3. Iodine Uptake and Release

To prevent the presence of radioactive toxic gases in the environment, researchers have focused on the preparation of selective porous materials. As medi-MOF-1 has developed porosity and a stable crystalline structure, the I2 uptake experiment was carried out in the vapor phase. Before the adsorption experiments were started, the sample was activated at 60 °C for 6 h under a vacuum. Iodine uptake was measured using the gravimetric method. In the vapor phase, 30 mg of medi-MOF-1 [39] was placed in an I2 chamber at 298 K. After 12 h, the color of medi-MOF-1 changed from orange-red to dark brown. No further change was observed after 10 h, and the maximum adsorbed amount of I2 was as high as 1.936 g g−1 (Figure 4a). Compared to other typically porous MOFs usually used for I2 adsorption via vapor diffusion, medi-MOF-1 also exhibits higher I2 adsorption. The relevant BET data and iodine adsorption capacities of the selected MOFs are summarized and tabulated in Table 2. This result may be attributed to the frameworks with conjugated π-electrons, which could produce multiple interactions for iodine [54,55]. Iodine templates are introduced into the assembly process of MOFs such as (ZnI2)3(tpt)2 at room temperature in order to affect the uptake capacities of iodine [56]. Owing to medi-MOF-1 can keep its crystal structure unchanged in ethanol, we soaked 100 mg of medi-MOF-1 crystals in 3 mL of a dry ethanol solution of I2 in a sealed glass vial at room temperature. After 48 h, I2 molecules were mostly adsorbed by the free active sites in medi-MOF-1, and no more free sites were left. There is almost no N2 sorption in the low-pressure region after the incorporation of iodine, indicating that I2 completely fills the pores (Figure 4b). A PXRD study was carried out before and after the I2 absorption experiment and proved that the framework of medi-MOF-1 retained the host framework crystallinity after loading the I2 molecules (Figure 4c). Thermogravimetric analysis (TGA) was performed to check the I2 loading amount. The I2@MOF showed a weight loss of ~52% from 100 to 500 °C (Figure 4d). Combined with the thermogravimetric curve and the molecular formula of medi-MOF-1, it can be calculated as 500 mg of iodine per gram of MOF. Compared to other porous MOFs usually used for I2 adsorption via solution-based processes in ethanol, medi-MOF-1 also exhibits higher iodine adsorption than JLU-Liu14 [57,58]. The high loading of I2 for medi-MOF-1 can be attributed to the different pore sizes of these MOFs and the possible interaction between the porous skeleton and I2.
From the point of view of recyclability, the I2 desorption from the porous framework is also essential. The I2 release process was detected by UV-visible (UV-vis) spectra. The captured I2 could be easily separated from the frameworks upon immersion in I2-loaded medi-MOF-1 in ethanol. A UV-vis spectrophotometer recorded the release of I2 from the medi-MOF-1 framework at different times at room temperature. When the I2@medi-MOF-1 was soaked in dry ethanol, the color of the iodine-loaded sample changed gradually from dark brown to orange-red. Afterwards, the release slows down and subsequently, the color of the ethanol solution changes from colorless to yellow. As illustrated in Figure 5, UV-vis spectra show absorption bands at λmax = 218 and 263 nm, which can be attributed to I2. And the band observed at 263 nm may be assigned to polyiodide ions (I3), established due to the reaction of I2 with decomposed iodide. The release of iodine increased with time, suggesting that this desorption behavior of iodine is based on host–guest interactions. This adsorption and release of I2 by medi-MOF-1 reveals that the I2 sorption process of the MOF is reversible.

3. Materials and Methods

3.1. Materials and General Methods

All reagents were obtained from commercial sources and used as received. All the other chemical reagents used were of AR grade. PXRD was collected on a Rigaku D/Max 2550 X-ray diffractometer with Cu-Kα radiation (λ = 1.5418 Å). FT-IR spectra were obtained by a Nicolet Impact 410 Fourier-transform infrared spectrometer in the 400–4000 cm−1 range with KBr pellets. TGA was performed under an air atmosphere in the range of 30–800 °C at a heating rate of 10 °C min−1 using a Perkin-Elmer TGA 7 thermogravimetric analyzer. Gas adsorption and desorption isotherms were measured on Quantachrom Autosorb-iQ after degassing the sample for 8 h at 100 °C. H2 adsorption tests were performed at 77 and 87 K. CO2, CH4, and N2 adsorption tests were performed at 273 K. The morphologies of the powder materials were recorded using a JEOL-JSM-6700F field-emission scanning electron microscope (SEM).

3.2. Synthesis of Medi-MOF-1 and Iodine-Loaded Medi-MOF-1

Medi-MOF-1 have been produced on a large scale. A mixture of Zn(OAc)2·2H2O (200 mg, 0.9112 mmol), curcumin (600 mg, 1.6287 mmol), N,N-dimethylformamide (40 mL), and ethanol (10 mL) was sealed into a 100 mL capped vessel. The vessel was heated at 75 °C. The crystals were obtained after 5 days and washed with DMF. Yield: 55% (based on curcumin). Iodine uptake experiment was carried out via vapor diffusion and solution phase. A 100 mg sample was immersed in an ethanol solution of iodine. The complete absorption experiment was done at room temperature for 48 h.

4. Conclusions

In summary, we report on large-scale reactions to synthesize medi-MOF-1 toward the capture of I2 and gas sorption. We utilized medi-MOF-1 based on curcumin, which displays an interesting and important ability to capture I2 (1.936 g g−1) at room temperature. The medi-MOF-1 microcrystals have lower Langmuir specific surface areas than previously published medi-MOF-1. In particular, it is amongst the more efficient porous materials for I2 adsorption. Furthermore, medi-MOF-1 also exhibits hydrogen adsorption capacities of 1.57 wt.% at 77 K and 1 bar. The adsorption capacity of CO2 by activated medi-MOF-1 is greater than that of CH4 and N2 at 273 K, and the selectivities of CO2/N2 and CO2/CH4 are 5.8 and 2.5, respectively. This study aims to provide material for the potential application of I2 adsorption as well as gas adsorption and separations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28135237/s1, Figure S1: The fitting data for calculating the H2 Qst value for medi-MOF-1; Figure S2: FT-IR of medi-MOF-1 and medi-MOF-1@I2; Figure S3: Up: visual color change of iodine solution in ethanol I2 adsorption progress of medi-MOF-1, Below: Photographs showing the color change iodine capture for medi-MOF-1; Table S1: Summary of H2 uptake and Qst value for some reported MOFs at 77K, 1bar.

Author Contributions

Conceptualization, H.S.; methodology, Y.Z.; data curation, T.H.; writing—original draft, H.S.; writing—review & editing, H.S. and F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the financial support of the National Natural Science Foundation of China (21871105). This research was also funded by the financial support of the Natural Science Foundation of the Jiangsu Higher Education Institutions, China (21KJD150001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Md, M.R.; Abayomi, O.O.; Eskinder, G.; Amit, K. Assessment of energy storage technologies: A review. Energy Convers. Manag. 2020, 223, 113295. [Google Scholar]
  2. Rivera, F.P.; Zalamea, J.; Espinoza, J.L.; Gonzalez, L.G. Sustainable use of spilled turbinable energy in Ecuador: Three different energy storage systems. Renew. Sustain. Energy Rev. 2022, 156, 112005. [Google Scholar] [CrossRef]
  3. Denning, S.; Majid, A.A.A.; Crawford, J.M.; Carreon, M.A.; Koh, C.A. Promoting Methane Hydrate Formation for Natural Gas Storage over Chabazite Zeolites. ACS Appl. Energy Mater. 2021, 4, 13420–13424. [Google Scholar] [CrossRef]
  4. Kumar, S.; Bera, R.; Das, N.; Koh, J. Chitosan-based zeolite-Y and ZSM-5 porous biocomposites for H2 and CO2 storage. Carbohydr. Polym. 2020, 232, 115808. [Google Scholar] [CrossRef] [PubMed]
  5. Yue, B.; Liu, S.S.; Chai, Y.C.; Wu, G.J.; Guan, N.J.; Li, L.D. Zeolites for separation: Fundamental and application. J. Energy Chem. 2022, 71, 288–303. [Google Scholar] [CrossRef]
  6. Liu, S.R.; Chen, H.T.; Lv, H.X.; Qin, Q.P.; Fan, L.M.; Zhang, X.T. Chemorobust 4p-5p {InPb}-organic framework for efficiently catalyzing cycload6dition of CO2 with epoxides and deacetalization-Knoevenagel condensation. Mater. Today Chem. 2022, 24, 100984. [Google Scholar] [CrossRef]
  7. Gan, L.; Andres-Garcia, E.; Espallargas, G.M.; Planas, J.G. Adsorptive separation of CO2 by a hydrophobic carborane-based Metal−Organic Framework under humid conditions. ACS Appl. Mater. Interfaces 2023, 15, 5309–5316. [Google Scholar] [CrossRef]
  8. Zhang, T.; Chen, H.T.; Liu, S.R.; Lv, H.X.; Zhang, X.T.; Li, Q.L. Highly Robust {Ln4}-Organic Frameworks (Ln = Ho, Yb) for Excellent Catalytic Performance on Cycloaddition Reaction of Epoxides with CO2 and Knoevenagel Condensation. ACS Catal. 2021, 11, 14916–14925. [Google Scholar] [CrossRef]
  9. Guan, Q.; Zhou, L.L.; Dong, Y.B. Metalated covalent organic frameworks: From synthetic strategies to diverse applications. Chem. Soc. Rev. 2022, 51, 6307–6416. [Google Scholar] [CrossRef] [PubMed]
  10. Tian, Y.; Zhu, G.S. Porous aromatic frameworks. Chem. Rev. 2020, 120, 8934–8986. [Google Scholar] [CrossRef] [PubMed]
  11. Zhou, H.C.; Kitigawa, S. Metal-organic frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef] [Green Version]
  12. He, H.; Sun, Q.; Gao, W.; Perman, J.A.; Sun, F.; Zhu, G.; Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable Metal-Organic Framework Featuring a Local Buffer Environment for Carbon Dioxide Fixation. Angew. Chem. Int. Ed. 2018, 57, 4657–4662. [Google Scholar] [CrossRef]
  13. Liu, X.R.; Chen, H.T.; Zhang, X.T. Bifunctional {Pb10K2}–Organic Framework for High Catalytic Activity in Cycloaddition of CO2 with Epoxides and Knoevenagel Condensation. ACS Catal. 2022, 12, 10373–10383. [Google Scholar] [CrossRef]
  14. Li, X.F.; Bian, H.; Huang, W.Q.; Yan, B.Y.; Wang, X.Y.; Zhu, B. A review on anion-pillared metal-organic frameworks (APMOFs) and their composites with the balance of adsorption capacity and separation selectivity for efficient gas separation. Coord. Chem. Rev. 2022, 470, 214714. [Google Scholar] [CrossRef]
  15. He, H.M.; Wen, H.M.; Li, H.K.; Zhang, H.W. Recent advances in metal-organic frameworks and their derivatives for electrocatalytic nitrogen reduction to ammonia. Coord. Chem. Rev. 2022, 471, 214761. [Google Scholar] [CrossRef]
  16. Yuan, R.R.; Yan, Z.J.; Shaga, A.; He, H.M. Design and fabrication of an electrochemical sensing platform based on a porous organic polymer for ultrasensitive ampicillin detection. Sens. Actuator B Chem. 2021, 327, 128949. [Google Scholar] [CrossRef]
  17. Zhang, H.W.; Zhu, Q.Q.; Yuan, R.; He, H. Crystal Engineering of MOF@COF Core-Shell Composites for Ultra-Sensitively Electrochemical Detection. Sens. Actuators B Chem. 2021, 329, 129144. [Google Scholar] [CrossRef]
  18. He, H.; Zhu, Q.Q.; Li, C.P.; Du, M. Design of a Highly-Stable Pillar-Layer Zinc(II) Porous Framework for Rapid, Reversible, and Multi-Responsive Luminescent Sensor in Water. Cryst. Growth Des. 2019, 19, 694–703. [Google Scholar] [CrossRef]
  19. Fischer, M.; Hoffmann, F.; Froba, M. Preferred Hydrogen Adsorption Sites in Various MOFs-A Comparative Computational Study. ChemPhysChem 2009, 10, 2647–2657. [Google Scholar] [CrossRef]
  20. Suh, M.P.; Park, H.J.; Prasad, T.K.; Lim, D.W. Hydrogen Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 782–835. [Google Scholar] [CrossRef] [PubMed]
  21. Lamiel, C.; Hussain, I.; Rabiee, H.; Ogunsakin, O.R.; Zhang, K.L. Metal-organic framework-derived transition metal chalcogenides (S, Se, and Te): Challenges, recent progress, and future directions in electrochemical energy storage and conversion systems. Coord. Chem. Rev. 2023, 480, 215030. [Google Scholar] [CrossRef]
  22. Phan, A.; Doonan, C.J.; Uribe-Romo, F.J.; Knobler, C.B.; O’Keeffe, M.; Yaghi, O.M. Synthesis, Structure, and Carbon Dioxide Capture Properties of Zeolitic Imidazolate Frameworks. Acc. Chem. Res. 2010, 43, 58–67. [Google Scholar] [CrossRef] [PubMed]
  23. He, H.; Perman, J.A.; Zhu, G.; Ma, S. Metal-Organic Frameworks for CO2 Chemical Transformations. Small 2016, 12, 6309–6324. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, H.; Zhang, T.; Liu, S.; Lv, H.; Fan, L.; Zhang, X. Fluorine-Functionalized NbO-Type {Cu2}-Organic Framework: Enhanced Catalytic Performance on the Cycloaddition Reaction of CO2 with Epoxides and Deacetalization-Knoevenagel Condensation. Inorg. Chem. 2022, 61, 11949–11958. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, H.T.; Liu, S.R.; Lv, H.X.; Qin, Q.P.; Zhang, X.T. Nanoporous {Y2}-Organic Frameworks for Excellent Catalytic Performance on the Cycloaddition Reaction of Epoxides with CO2 and Deacetalization–Knoevenagel Condensation. ACS Appl. Mater. Interfaces 2022, 14, 18589–18599. [Google Scholar] [CrossRef]
  26. Xu, G.R.; An, Z.H.; Xu, K.; Liu, Q.; Das, R.; Zhao, H.L. Metal organic framework (MOF)-based micro/nanoscaled materials for heavy metal ions removal: The cutting-edge study on designs, synthesis, and applications. Coord. Chem. Rev. 2021, 427, 213554. [Google Scholar] [CrossRef]
  27. Ma, L.; Huang, C.; Yao, Y.Y.; Fu, M.T.; Han, F.; Li, Q.N.; Wu, M.H.; Zhang, H.J.; Xu, L.; Ma, H.J. Self-assembled MOF microspheres with hierarchical porous structure for efficient uranium adsorption. Sep. Purif. Technol. 2023, 314, 123526. [Google Scholar] [CrossRef]
  28. Zhang, X.R.; Maddock, J.; Nenoff, T.M.; Denecke, M.A.; Yang, S.H.; Schroer, M. Adsorption of iodine in metal-organic framework materials. Chem. Soc. Rev. 2022, 51, 3243–3262. [Google Scholar] [CrossRef]
  29. Tang, Y.Z.; Huang, H.L.; Li, J.; Xue, W.J.; Zhong, C.L. IL-induced formation of dynamic complex-iodine anions in IL@MOF composites for efficient iodine capture. J. Mater. Chem. A 2019, 7, 18324–18329. [Google Scholar] [CrossRef]
  30. Mandal, A.; Adhikary, A.; Sarkar, A.; Das, D. Naked Eye Cd2+ Ion Detection and Reversible Iodine Uptake by a Three-Dimensional Pillared-Layered Zn-MOF. Inorg. Chem. 2020, 59, 17758–17765. [Google Scholar] [CrossRef]
  31. Sarkar, A.; Adhikary, A.; Mandal, A.; Chakraborty, T.; Das, D. Zn-BTC MOF as an Adsorbent for Iodine Uptake and Organic Dye Degradation. Cryst. Growth Des. 2020, 20, 7833–7839. [Google Scholar] [CrossRef]
  32. Lee, J.W.; Cha, D.K.; Oh, Y.K.; Ko, K.B.; Song, J.S. Zero-valent iron pretreatment for detoxifying iodine in liquid crystal display (LCD) manufacturing wastewater. J. Hazard. Mater. 2009, 164, 67–72. [Google Scholar] [CrossRef] [PubMed]
  33. Chapman, K.W.; Chupas, P.J.; Nenoff, T.M. Radioactive Iodine Capture in Silver-Containing Mordenites through Nanoscale Silver Iodide Formation. J. Am. Chem. Soc. 2010, 132, 8897–8899. [Google Scholar] [CrossRef] [PubMed]
  34. Riebe, B.; Dultz, S.; Bunnenberg, C. Temperature effects on iodine adsorption on organo-clay minerals: I. Influence of pretreatment and adsorption temperature. Appl. Clay Sci. 2005, 28, 9–16. [Google Scholar] [CrossRef]
  35. 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. Horiz. 2019, 6, 1571–1595. [Google Scholar] [CrossRef]
  36. Zeng, M.H.; Wang, Q.X.; Tan, Y.X.; Hu, S.; Zhao, H.X.; Long, L.S.; Kurmoo, M. Rigid Pillars and Double Walls in a Porous Metal-Organic Framework: Single-Crystal to Single-Crystal, Controlled Uptake and Release of Iodine and Electrical Conductivity. J. Am. Chem. Soc. 2010, 132, 2561–2563. [Google Scholar] [CrossRef]
  37. Wang, C.; Tian, L.; Zhu, W.; Wang, P.; Liang, Y.; Zhang, W.L.; Zhao, H.W.; Li, G.T. Dye@bio-MOF-1 Composite as a Dual-Emitting Platform for Enhanced Detection of a Wide Range of Explosive Molecules. ACS Appl. Mater. Interfaces 2017, 9, 20076–20085. [Google Scholar] [CrossRef]
  38. Yan, Z.; Qiao, Y.; Wang, J.; Xie, J.; Cui, B.; Fu, Y.; Lu, J.; Yang, Y.; Bu, N.; Yuan, Y.; et al. An Azo-Group-Functionalized Porous Aromatic Framework for Achieving Highly Efficient Capture of Iodine. Molecules 2022, 27, 6297. [Google Scholar] [CrossRef]
  39. Kamal, S.; Khalid, M.; Khan, M.S.; Shanid, M.; Ahmad, M. Amine- and Imine-Functionalized Mn-Based MOF as an Unusual Turn-On and Turn-Off Sensor for d10 Heavy Metal Ions and an Efficient Adsorbent to Capture Iodine. Cryst. Growth Des. 2022, 22, 3277–3294. [Google Scholar] [CrossRef]
  40. Sava, D.F.; Chapman, K.W.; Rodriguez, M.A.; Greathouse, J.A.; Crozier, P.S.; Zhao, H.Y.; Chupas, P.J.; Nenoff, T.M. Competitive I2 Sorption by Cu-BTC from Humid Gas Streams. Chem. Mater. 2013, 25, 2591–2596. [Google Scholar] [CrossRef]
  41. Marshall, R.J.; Griffin, S.L.; Wilson, C.; Forgan, R.S. Single-Crystal to Single-Crystal Mechanical Contraction of Metal-Organic Frameworks through Stereoselective Postsynthetic Bromination. J. Am. Chem. Soc. 2015, 137, 9527–9530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Su, H.M.; Sun, F.X.; Jia, J.T.; He, H.M.; Wang, A.F.; Zhu, G.S. A highly porous medical metal-organic framework constructed from bioactive curcumin. Chem. Commun. 2015, 51, 5774–5777. [Google Scholar] [CrossRef] [PubMed]
  43. Feng, X.D.; Wang, Y.F.; Muhammad, F.; Sun, F.X.; Tian, Y.Y.; Zhu, G.S. Size, Shape, and Porosity Control of Medi-MOF-1 via Growth Modulation under Microwave Heating. Cryst. Growth Des. 2019, 19, 889–895. [Google Scholar] [CrossRef]
  44. He, H.; Xue, Y.Q.; Wang, S.Q.; Zhu, Q.Q.; Chen, J.; Li, C.P. A Double-Walled Bimetal-Organic Framework for Antibiotics Sensing and Size-Selective Catalysis. Inorg. Chem. 2018, 57, 15062–15068. [Google Scholar] [CrossRef]
  45. An, J.; Geib, S.J.; Rosi, N.L. High and Selective CO2 Uptake in a Cobalt Adeninate Metal-Organic Framework Exhibiting Pyrimidine- and Amino-Decorated Pores. J. Am. Chem. Soc. 2010, 132, 38–39. [Google Scholar] [CrossRef] [PubMed]
  46. Freund, R.; Zaremba, O.; Dinca, M.; Arnauts, G.; Ameloot, R.; Skorupskii, G.; Bavykina, A.; Gascon, J.; Ejsmont, A.; Goscianska, J.; et al. The Current Status of MOF and COF Applications. Angew. Chem. Int. Ed. 2021, 60, 23975–24001. [Google Scholar] [CrossRef]
  47. Chen, Y.F.; Jiang, J.W. A Bio-Metal-Organic Framework for Highly Selective CO2 Capture: A Molecular Simulation Study. ChemSusChem 2010, 3, 982–988. [Google Scholar] [CrossRef]
  48. Levine, D.J.; Runcevski, T.; Kapelewski, M.T.; Keitz, B.K.; Oktawiec, J.; Reed, D.A.; Mason, J.A.; Jiang, H.Z.H.; Colwell, K.A.; Legendre, C.M.; et al. Olsalazine-Based Metal-Organic Frameworks as Biocompatible Platforms for H2 Adsorption and Drug Delivery. J. Am. Chem. Soc. 2016, 138, 10143–10150. [Google Scholar] [CrossRef] [Green Version]
  49. An, J.; Rosi, N.L. Tuning MOF CO2 Adsorption Properties via Cation Exchange. J. Am. Chem. Soc. 2010, 132, 5578–5579. [Google Scholar] [CrossRef]
  50. Zulys, A.; Yulia, F.; Muhadzib, N.; Nasruddin. Biological Metal-Organic Frameworks (Bio-MOFs) for CO2 Capture. Ind. Eng. Chem. Res. 2021, 60, 37–51. [Google Scholar] [CrossRef]
  51. Han, B.; Chakraborty, A.; Saha, B.B. Isosteric Heats and Entropy of Adsorption in Henry’s Law Region for Carbon and MOFs Structures for Energy Conversion Applications. Int. J. Heat Mass Transf. 2022, 182, 122000. [Google Scholar] [CrossRef]
  52. Rojas, S.; Devic, T.; Horcajada, P. Metal organic frameworks based on bioactive components. J. Mater. Chem. B 2017, 5, 2560–2573. [Google Scholar] [CrossRef] [PubMed]
  53. Li, T.; Chen, D.L.; Sullivan, J.E.; Kozlowski, M.T.; Johnson, J.K.; Rosi, N.L. Systematic modulation and enhancement of CO2: N2 selectivity and water stability in an isoreticular series of bio-MOF-11 analogues. Chem. Sci. 2013, 4, 1746–1755. [Google Scholar] [CrossRef]
  54. Safarifard, V.; Morsali, A. Influence of an amine group on the highly efficient reversible adsorption of iodine in two novel isoreticular interpenetrated pillared-layer microporous metal-organic frameworks. Crystengcomm 2014, 16, 8660–8663. [Google Scholar] [CrossRef]
  55. Yan, Z.J.; Yuan, Y.; Tian, Y.Y.; Zhang, D.M.; Zhu, G.S. Highly Efficient Enrichment of Volatile Iodine by Charged Porous Aromatic Frameworks with Three Sorption Sites. Angew. Chem. Int. Ed. 2015, 54, 12733–12737. [Google Scholar] [CrossRef] [PubMed]
  56. Gee, W.J.; Hatcher, L.E.; Cameron, C.A.; Stubbs, C.; Warren, M.R.; Burrows, A.D.; Raithby, P.R. Evaluating Iodine Uptake in a Crystalline Sponge Using Dynamic X-ray Crystallography. Inorg. Chem. 2018, 57, 4959–4965. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, J.; Luo, J.H.; Luo, X.L.; Zhao, J.; Li, D.S.; Li, G.H.; Huo, Q.S.; Liu, Y.L. Assembly of a Three-Dimensional Metal-Organic Framework with Copper(I) Iodide and 4-(Pyrimidin-5-yl) Benzoic Acid: Controlled Uptake and Release of Iodine. Cryst. Growth Des. 2015, 15, 915–920. [Google Scholar] [CrossRef]
  58. Yang, Y.T.; Tu, C.Z.; Yin, H.J.; Liu, J.J.; Cheng, F.X.; Luo, F. Molecular Iodine Capture by Covalent Organic Frameworks. Molecules 2022, 27, 9045. [Google Scholar] [CrossRef]
  59. Marshall, R.J.; Griffin, S.L.; Wilson, C.; Forgan, R.S. Stereoselective Halogenation of Integral Unsaturated C-C Bonds in Chemically and Mechanically Robust Zr and Hf MOFs. Chem.-Eur. J. 2016, 22, 4870–4877. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. The synthetic process of medi-MOF-1.
Scheme 1. The synthetic process of medi-MOF-1.
Molecules 28 05237 sch001
Figure 1. (a) Trinuclear Zn(II) clusters of medi-MOF-1 (Zn, green; C, gray; O, red; H are omitted for clarity.); (b) PXRD patterns of medi-MOF-1 (black, simulated; red, as-synthesized medi-MOF-1); (c,d) SEM images of the as-synthesized medi-MOF-1.
Figure 1. (a) Trinuclear Zn(II) clusters of medi-MOF-1 (Zn, green; C, gray; O, red; H are omitted for clarity.); (b) PXRD patterns of medi-MOF-1 (black, simulated; red, as-synthesized medi-MOF-1); (c,d) SEM images of the as-synthesized medi-MOF-1.
Molecules 28 05237 g001
Figure 2. (a) Hydrogen adsorption and desorption isotherms of medi-MOF-1 (red circle at 77 K, blue triangle at 87 K); (b) the calculated Qst plots for H2 uptake in the medi-MOF-1.
Figure 2. (a) Hydrogen adsorption and desorption isotherms of medi-MOF-1 (red circle at 77 K, blue triangle at 87 K); (b) the calculated Qst plots for H2 uptake in the medi-MOF-1.
Molecules 28 05237 g002
Figure 3. (a) CO2, CH4, and N2 adsorption isotherms at 273 K for medi-MOF-1; (b) CO2/CH4/N2 selectivity for medi-MOF-1 (273 K) calculated using the Henry’s law constants in the linear low pressure (<0.1 bar) range.
Figure 3. (a) CO2, CH4, and N2 adsorption isotherms at 273 K for medi-MOF-1; (b) CO2/CH4/N2 selectivity for medi-MOF-1 (273 K) calculated using the Henry’s law constants in the linear low pressure (<0.1 bar) range.
Molecules 28 05237 g003
Figure 4. (a) Iodine vapor adsorption curves at 298 K under ambient pressure for medi-MOF-1; (b) N2 adsorption and desorption isotherms of activated medi-MOF-1 (red), I2@medi-MOF-1 (blue).; (c) PXRD patterns of activated medi-MOF-1(black), and I2@medi-MOF-1 (red); (d) the TGA curve of I2@medi-MOF-1.
Figure 4. (a) Iodine vapor adsorption curves at 298 K under ambient pressure for medi-MOF-1; (b) N2 adsorption and desorption isotherms of activated medi-MOF-1 (red), I2@medi-MOF-1 (blue).; (c) PXRD patterns of activated medi-MOF-1(black), and I2@medi-MOF-1 (red); (d) the TGA curve of I2@medi-MOF-1.
Molecules 28 05237 g004
Figure 5. Up: Photographs of the time-dependent I2 desorption process by medi-MOF-1 in ethanol; below: UV-vis spectra for I2 release from medi-MOF-1 in ethanol.
Figure 5. Up: Photographs of the time-dependent I2 desorption process by medi-MOF-1 in ethanol; below: UV-vis spectra for I2 release from medi-MOF-1 in ethanol.
Molecules 28 05237 g005
Table 1. CO2 adsorption data of MOFs for some reported bio-MOFs at 77 K, 1 bar.
Table 1. CO2 adsorption data of MOFs for some reported bio-MOFs at 77 K, 1 bar.
MaterialsBET
(m2 g−1)
Pore Volume
(c g−1)
CO2 @273 K
(cm3 g−1)
Ref.
bio-MOF-116800.7576[49]
bio-MOF-1111480.45147[53]
bio-MOF-1210080.42100[53]
bio-MOF-134120.2060[53]
bio-MOF-14170.03545[53]
medi-MOF-1475-34.7This work
Table 2. Iodine adsorption capacities in MOFs via vapor diffusion.
Table 2. Iodine adsorption capacities in MOFs via vapor diffusion.
MaterialsBET
(m2 g−1)
Pore Volume
(cm3 g−1)
Iodine Uptake
(wt.%)
Ref.
Cu-BTC18500.74175[40]
ZIF-816300.66125[54]
(ZnI2)3(tpt)2--173[56]
Zr6O4(OH)4(sdc)629001.33107[59]
Zr6O4(OH)4(peb)626501.16279[59]
medi-MOF-1475-193This 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

Su, H.; Zhou, Y.; Huang, T.; Sun, F. Study on Gas Sorption and Iodine Uptake of a Metal-Organic Framework Based on Curcumin. Molecules 2023, 28, 5237. https://doi.org/10.3390/molecules28135237

AMA Style

Su H, Zhou Y, Huang T, Sun F. Study on Gas Sorption and Iodine Uptake of a Metal-Organic Framework Based on Curcumin. Molecules. 2023; 28(13):5237. https://doi.org/10.3390/molecules28135237

Chicago/Turabian Style

Su, Hongmin, Yang Zhou, Tao Huang, and Fuxing Sun. 2023. "Study on Gas Sorption and Iodine Uptake of a Metal-Organic Framework Based on Curcumin" Molecules 28, no. 13: 5237. https://doi.org/10.3390/molecules28135237

Article Metrics

Back to TopTop