Abstract

The preparation of a functional device based on a functionalized MIL-100(Fe) metal–organic framework for the solid-phase extraction of heavy metals is reported. By a simple and easy straightforward grafting procedure, a thiol-functionalized MIL-100(Fe) material (MIL-100(Fe)-SH) with a S/Fe ratio of 0.80 and a surface area of 840 m2 g–1 was obtained. MIL-100(Fe)-SH exhibited a higher Hg(II) extraction (96 ± 5%) than that of MIL-100(Fe) (78 ± 4%) due to the interaction between thiol groups and Hg(II) ions. For practical applications, the obtained MIL-100(Fe)-SH was integrated by a simple method to a 3D printed support based on a matrix of interconnected cubes using poly(vinylidene fluoride) as binder, obtaining a functional device that simultaneously acts as stirrer and sorbent. The developed device showed high efficiency for the removal of Hg(II), good reusability, and excellent performance for the simultaneous preconcentration and further detection and quantification of Hg(II), Pb(II), and As(V) in tap, well, and lake water samples.
Short abstract
The functional MIL-100(Fe)-SH/device, which simultaneously acts as stirrer and sorbent, showed excellent performance for the solid-phase extraction of Hg(II), Pb(II), and As(V) in real water samples.
1. Introduction
Access to safe and clean potable water is essential for humans and wildlife. Water pollution with a variety of different pollutants, including pathogens, persistent organic pollutants, and heavy metals, has become a global environmental problem.1 Among them, heavy metals, which are metals and metalloids with a relative density equal or superior to 5 g cm–3,2 are a cause of growing concern due to their high toxicity and adverse effects on aquatic biota and human health even at low levels.3 Because of their continuous release into the environment from different natural and anthropogenic sources and their persistent and nonbiodegradable nature, different heavy metals have been detected in groundwater, seawater and wastewater.4−6 Several traditional water treatments, such as chemical precipitation,7 chlorination,8 ion exchange,9 and membrane filtration,10 have been used to remove heavy metals from water. However, some of these methods are ineffective or expensive or generate toxic residues. In this context, the use of porous solids as sorbents has emerged as a simple and effective alternative in the removal of heavy metals from water.11−13 Owing to their unique characteristics, such as outstanding porosity, extreme versatility and tunability, thermal and chemical stability, and abundant active sites, metal–organic frameworks (MOFs) have shown great potential as advanced sorbents of different pollutants, including heavy metals.14−16 Between the different types of MOFs synthesized so far, the Materials of Institute Lavoisier frameworks (MILs), composed of trivalent metal ions and carboxylic acid groups, are among the most studied for these and other applications thanks to their excellent surface area and water stability.17
It has been reported that the incorporation of functional groups on MOFs via functionalization of their linkers and/or metal nodes is a promising approach to improve their extraction capacity of heavy metals.18 For instance, an amino-functionalized MIL-101(Cr) was prepared through the coordination of the open Cr3+ sites with ethylenediamine and used for Pb2+removal.19 The Pb2+ adsorption capacity of the obtained material was 81 mg g–1, which was more than 5 times that of MIL-101 without amino groups. Following a similar approach, a sulfonic acid grafted Cu-MOF, Cu3(BTC)2-SO3H, was obtained and used for the elimination of Cd2+ ions, obtaining a high cadmium uptake capacity of 88.7 mg g–1.20 Using 2-aminoterephthalic acid as a ligand, an NH2-MIL-53(Al) MOF was also synthesized and tested for the adsorption of Hg2+, showing excellent extraction capacity due to the strong interaction between −NH2 groups and Hg2+ ions.21 More recently, polyethyleneimine modified UiO-66-NH2 nanoparticles with excellent Pb2+ adsorption capacity, which was attributed to the large amount of −NH2 and −NH– groups of polyethyleneimine molecules, were reported.22 The potential use of thiol-functionalized HKUST-1 and UiO-66 as sorbents for the removal of Hg2+ ions have been also demonstrated.23
However, the powder nature of MOF materials limits its applicability for the removal of pollutants from water because tedious and lengthy procedures are necessary to separate the sorbent from the treated water. To overcome this problem and improve the performance of MOFs, they have been immobilized on different supports such as membranes,24 monoliths,25 sponges,26 and fibers.27 In this context, 3D printing technology has appeared as an interesting and useful alternative for the fabrication of supports with complex structures in short periods of time.28−30 Moreover, not long ago, the combination of MOFs with 3D printing technology has allowed the development of different functional devices for water treatment.31−34
In this work, we report the preparation of thiol-functionalized MOF-coated 3D printed devices for the removal of heavy metals from wastewaters. The water-stable and ecofriendly MIL-100(Fe) MOF was functionalized by an easy grafting process, obtaining a thiol-functionalized material (MIL-100(Fe)-SH). The MIL-100(Fe)-SH MOF was applied to the removal of mercury from water, and the performance was compared with the nonfunctionalized MIL-100(Fe). To improve its applicability as an adsorbent, the obtained material was incorporated on a 3D printed device by a fast-coating procedure, obtaining a functional device that simultaneously acts as stirrer and sorbent. The influence of pH on the sample solution and contact time on the removal of Hg(II) was studied. In addition, the potential of the developed device for simultaneous removal of Hg(II), Pb(II), and As(V) from water samples was also evaluated.
2. Experimental Section
2.1. Instrumentation
XRD diffraction and XRD microdiffraction patterns were acquired utilizing Cu Kα radiation on Siemens D5000 and Bruker D8 diffractometers, respectively. A simulated pattern of bulk MIL-100 was obtained from the crystallographic data reported in the literature35 using the Mercury V.3.10.3 software. Thermogravimetric analysis (TGA) was carried out in a nitrogen atmosphere using a TA Instruments SDT 2960 simultaneous DSC-TGA. N2 adsorption–desorption isotherms were acquired at 77 K using a TriStar II (Micromeritics) gas adsorption instrument. The samples were previously outgassed under a dynamic vacuum at 423 K for 12 h. Scanning electron microscopy (Hitachi S-3400N) equipped with a Bruker AXS Xflash 4010 EDS system was used to study the morphology and elemental composition of the samples. Infrared spectroscopy (FTIR) using CO as a probe molecule was performed with a Bruker Vertex 80v spectrophotometer equipped with an MCT detector operating at 3 cm–1 resolution. For that, the samples were prepared and activated inside the IR cell under a high vacuum at 423 K for 8 h.
2.2. Chemicals
All reagents were of analytical grade. Hydrochloric acid (HCl), potassium hexacyanoferrate(III) (K3[Fe(CN)6]), and potassium hydroxide (KOH) were purchased from J.T. Baker. Iron powder (Fe), poly(vinylidene fluoride) (PVDF, MW ∼180,000), potassium borohydride (KBH4), toluene, p-benzoquinone, and ammonium hydroxide (NH4OH) were acquired from Sigma Aldrich. Trimesic acid (H3BTC) and 1,2-ethanedithiol were obtained from Tokyo Chemical Industries. Stock solutions of 1000 mg L–1 of each studied metal, N,N-dimethylformamide (DMF), nitric acid (HNO3), and ethanol were purchased from Scharlab. The aqueous solutions used for the extraction experiments were prepared with ultrapure water through a Milli-Q water purification apparatus.
2.3. Synthesis of MIL-100(Fe)
Iron-based MIL-100 was prepared at room temperature following a simple procedure reported in the literature.36 Basically, a mixture of 0.28 g of Fe, 0.21 mL of HNO3, and 25 mL of deionized water was sonicated for 15 min. Later, 0.7 g of trimesic acid and 5.4 mg of p-benzoquinone were introduced, and the mixture was left under continuous stirring for 12 h. After this time, 12.5 mL of DMF was added and stirred for 2 h more. Lastly, the resulting orange solid was centrifuged and washed three times with ethanol.
2.4. Synthesis of MIL-100(Fe)-SH
The functionalization of MIL-100(Fe) with thiol groups was conducted following an adaptation of an experimental protocol described in a previous report.23 First, 0.4 g of MIL-100(Fe) was activated, inside a three-round-neck flask, at 423 K for 24 h under N2 flow to remove the solvent molecules coordinated to the iron centers. Then, 40 mL of anhydrous toluene and 6 mL of a 0.24 M 1,2 ethanedithiol/toluene solution were added, and the mixture was kept under stirring for 12 h at room temperature. The obtained powder was washed with ethanol to remove the unreacted 1,2-ethanedithiol molecules and dried under a vacuum.
2.5. Preparation of MIL-100(Fe)-SH/3D devices
The 3D device was designed using the Rhinoceros 6 software (McNeel & Associates, USA) and printed utilizing a Form 2 3D Printer (Formlabs) with Formlabs Clear V4 photoactive resin (FLGPCL04), which is characterized by low cost, smooth surface finish, and fine features and high detail. It consists of methacrylate monomers/oligomers and an initiator. The printing time was between 100 to 254 min to make 1 to 50 devices, respectively, with a 25 μm layer resolution and using 0.25 mL of liquid resin per device. Once printed, the devices were cleaned with 2-propanol to remove unreacted monomers and cured under UV radiation for 4 h. The immobilization of MIL-100(Fe)-SH on the 3D printed device was carried out following an easy coating method based on a previous published report.37 Basically, an acetone suspension of MIL-100(Fe)-SH (150 mg of MOF/5 mL acetone) was mixed with 1 g of DMF solution containing 7.5 wt % of poly(vinylidene fluoride) polymer by sonication. A stream of nitrogen gas was applied for acetone removal, and the resulting ink was added drop by drop on the 3D printed support, which was further dried at 333 K for 1 h.
2.6. Metal Extraction
The metal adsorption capacity of the developed materials and devices was studied by using 50 mL of a 1 mg L–1 heavy metal solution under continuous stirring. After extraction, the metal concentration in the supernatants was determined by ICP-OES (Optima 5300 DV, Perkin-Elmer) and CV/ HG-AFS (AFS-640, Rayleigh). The conditions are summarized in Tables S1 and S2, respectively.
3. Results and Discussion
3.1. Preparation of MIL-100(Fe)-SH
The thiol-functionalized MOF was obtained by simple grafting of 1,2-ethanedithiol on the open iron centers of the water-stable and highly porous MIL-100(Fe). The X-ray diffractograms of the as-synthesized MIL-100(Fe) and the thiol-grafted MIL-100(Fe) are shown in Figure 1a. Both materials showed high crystallinity, and the XRD patterns matched well with the simulated one, indicating the preservation of the structure after the grafting process. The incorporation of thiol groups on MIL-100(Fe) was demonstrated by EDS. In comparison with the EDS spectrum of MIL-100(Fe) (Figure 1b), an additional band at 2.3 KeV is observed in the spectrum of the functionalized MIL-100(Fe), which corresponds to the S (Kα) signal. The EDS analysis (Table S3) indicates an S/Fe ratio of 0.80. Furthermore, elemental EDS mapping of MIL-100(Fe)-SH (Figure 1c) shows a homogeneous distribution of S in the material. The functionalization of MIL-100(Fe) to yield MIL-100(Fe)-SH was evidenced by FTIR spectroscopy (Figure S1). The FTIR spectra of both samples show the characteristic IR peaks of MIL-100 MOF at 1626, 1447, and 1380 cm–1 that are assigned to the C–O stretching vibration of the carboxylic group and symmetric and asymmetric vibration of the OCO group, respectively.38 The two absorption bands around 2970 and 2886 cm–1 and the weak band around 2575 cm–1 observed in the spectrum of the MIL-100(Fe)-SH sample are attributed to C–H and S–H stretching vibrations of the 1,2 ethanedithiol molecules, respectively, indicating thiol functionalization of the MOF.39,40 The grafting of 1,2 ethanedithiol molecules to the open iron centers of MIL-100(Fe) was also studied by FTIR spectroscopy of adsorbed CO at 100 K. The IR spectra of CO adsorbed on activated MIL-100(Fe) before and after functionalization are shown in Figure 1d. Adsorption of CO on MIL-100(Fe) resulted in an asymmetric peak at 2170 cm–1, which, in agreement with other authors,41,42 is assigned to the stretching vibration of CO adsorbed on the open iron sites. This band is also present in the IR spectrum of CO adsorbed on the functionalized MIL-100(Fe)-SH, although of lower intensity, indicating that the coordinately unsaturated iron centers are partially occupied by the thiol molecules. The morphologies of the prepared samples were investigated by scanning electron microscopy. As can be observed in the corresponding SEM micrographs (Figure S2), both MOFs are formed by aggregates of particles with globular shape, indicating that the functionalization does not alter the morphology of the MOF. TGA analysis was carried out to evaluate the thermal stability of the grafted sample (Figure S3). The TGA curve of MIL-100(Fe)-SH showed a continuous weight loss, attributed to the removal of grafting 1,2-ethanedithiol and solvent molecules, followed by degradation of the framework at about 380 °C, which is similar to that of pristine MIL-100(Fe).
Figure 1.

(a) XRD patterns of MIL-100(Fe) and MIL-100(Fe)-SH. (b) EDS spectra of MIL-100(Fe) and MIL-100(Fe)-SH. (c) Fe and S EDS mappings of MIL-100(Fe)-SH. (d) FTIR spectra of CO adsorbed at 100 K on MIL-100(Fe) and MIL-100(Fe)-SH.
The BET surface area values of MIL-100(Fe) before and after functionalization, obtained from the analysis of the N2 isotherms, as well as the corresponding pore volume and pore size values are shown in Table 1. Both materials exhibit a high N2 uptake at a relative pressure lower than 0.1 (Figure 2a), indicating their microporous nature, which was confirmed by the pore size distributions (Figure 2b). However, a notable reduction of pore volume and surface area was produced after functionalization, which is due to thiol molecules partially occupying the space inside the pores. Similar results have been previously reported for the functionalization of different MOFs.43−45
Table 1. Sample Textural Analysis.
| sample | SBET (m2 g–1) | Vp (cm3 g–1) | pore width (Å) |
|---|---|---|---|
| MIL-100(Fe) | 1507 | 0.64 | 9–12/15–20 |
| MIL-100(Fe)-SH | 840 | 0.36 | 9–12/15–20 |
Figure 2.

(a) N2 adsorption–desorption isotherms and (b) pore-size curves of MIL-100(Fe) and MIL-100(Fe)-SH.
3.2. Adsorption of Mercury under Batch Conditions
The Hg extraction capacity of the MIL-100(Fe) and MIL-100(Fe)-SH MOFs was investigated under batch conditions. For that, 1 mg L–1 Hg(II) aqueous solution was put in contact with the prepared materials under continuous stirring during 24 h, and the removal capacity was determined using eq 1.
| 1 |
where C0 and Cf are the Hg concentrations before and after extraction. MIL-100(Fe) extracted 78 ± 4% of Hg(II), whereas MIL-100(Fe)-SH reached a 96 ± 5% extraction, demonstrating that, because of the strong soft acid–base interaction between −SH groups (soft base) and Hg (soft acid), the functionalization of the MOF with thiol groups improves the mercury adsorption capacity of the material.46−49 The adsorption mechanism was studied by measuring the pH values of the solution before and after the adsorption of Hg(II) by MIL-100(Fe)-SH (Figure S4). The pH of the solution becomes more acidic as the Hg(II) extraction increased, indicating the release of H+ ions and the probable ion exchange adsorption process during mercury uptake.40,50 In addition, it should be noted that the structure of MIL-100(Fe)-SH after extraction was not affected (Figure S5), indicating the high chemical stability of the functional material.
3.3. Characterization of MIL-100(Fe)-SH/3D Printed Devices
To improve the applicability of the developed adsorbent, it was incorporated on a 3D printed support by a facile coating procedure using an MIL-100(Fe)-SH/PVDF suspension. The 3D support is based on a matrix of interconnected cubes with a cylindrically shaped hole on the center of the device for the incorporation of a magnetic stirrer (Figure 3), which allows its simultaneous use as a stirrer and as sorbent.
Figure 3.
Design of the 3D printed support: (a) perspective view, (b) top view, and (c) front view.
Figure 4 shows the XRD patterns of the bare 3D printed support and after the coating with MIL-100(Fe)-SH. The diffractogram of the bare support does not show any diffraction line; however, the XRD diffractogram of the MIL-100(Fe)-SH/device exhibits the characteristic peaks of the MIL-100(Fe) structure, indicating that the applied coating procedure allows the incorporation of the MOF on the 3D support. The SEM images of the 3D support before and after the incorporation of MIL-100(Fe)-SH are shown in Figure 5. It can be observed that the uncoated support exhibited a smooth surface, which after the coating process appears covered by a uniform layer of MIL-100(Fe)-SH particles. Furthermore, sulfur (Figure 5e) and iron (Figure 5f) mappings showed the homogeneous distribution of both elements on the device, demonstrating the uniform deposition of MIL-100(Fe)-SH on the 3D support. The BET surface area of the MIL-100(Fe)-SH/device determined by N2 physisorption was 67 m2 g–1, which is given by the incorporation of the MOF on the support, as the bare device has zero surface area.
Figure 4.
XRD patterns of the 3D support before and after the coating procedure.
Figure 5.

SEM micrographs of the (a, b) bare 3D support and (c, d) MIL-100(Fe)-SH/device at different magnifications. (e) S and (f) Fe EDS mappings of the MIL-100(Fe)-SH/device.
3.4. Evaluation of the MIL-100(Fe)-SH/Device in the Extraction of Heavy Metals
The developed MIL-100(Fe)-SH/device was tested for the extraction of Hg(II). Figure 6a shows the effect of contact time on Hg(II) removal by the MIL-100(Fe)-SH/device. It can be seen that the removal percentage of Hg(II) increased by increasing the contact time from 2 h (38%) to 16 h (98%), and a further increment of the time did not improve it much. Therefore, 16 h was selected as the optimal extraction time. It should be noted that, at this time, the removal percentage of Hg(II) by a PVDF/device without the MOF was 37%, which indicates that the extraction capacity of the MIL-100(Fe)-SH/device is mainly due to the presence of the functionalized MOF.
Figure 6.

Effect on the Hg extraction capacity of (a) the contact time and (b) the sample pH (1 mg L–1, 50 mL, 24 h). (c) Recyclability of the MIL-100(Fe)/device for adsorption of Hg(II) from water.
In view of the influence of extraction medium pH on the stability of the adsorbent and on its extraction performance, the Fe leaching at the medium and the Hg(II) percentage removal at different pH values were studied. As shown in Figure 6b, the lowest Hg(II) adsorption was obtained at acid pH, which, according to the literature, may be due to the competition between the metal and the H+ ions for thiol groups.51−53 In addition, under these conditions, a small amount of iron was detected in the supernatant solution (Figure S6), indicating some degradation of the MIL-100(Fe)-SH material. As the pH of solution increased, the extraction percentage of Hg also increased, reaching the Hg(II) maximum adsorption capacity, with a negligible iron release in the solution at pH 6, which is probably due to the fact that, at this pH value, attractive electrostatic interaction can take place between the negatively charged surface of the MOF (Figure S7) and Hg(II). Thus, pH 6 was selected as the optimum pH value. To verify the potential reusability of the MIL-100(Fe)-SH/device, recycling tests were also performed. Between consecutive extractions, the device was washed with 0.1% thiourea–0.01 M HCl solution before reuse. As shown in Figure 6c, the extraction capacity of the MIL-100(Fe)-SH/device was almost identical after five Hg(II) adsorption–desorption cycles, which indicates the excellent reusability of the developed device.
To further evaluate the applicability and versatility of the developed MIL-100(Fe)-SH/device as an adsorbent, it was tested for the simultaneous extraction of Hg(II), Pb(II), and As(V) in tap, well, and lake water samples collected from Chihuahua (Mexico). The results are shown in Table 2. As can be observed, the obtained recoveries for Hg(II) and Pb(II) were 100% for all of them, whereas in the case of As(V), they ranged between 90 and 100% for well and tap water, respectively, which are comparable or even higher than the recoveries of these metals in real water samples reported using other adsorbents,54−57 demonstrating the potential of the developed device for the treatment of different water samples polluted with heavy metals. The low recovery of As(V) obtained for the lake water is probably due to the high amount of organic matter present in the sample, which limits the interaction of the metal with the thiol groups of MIL-100(Fe)-SH.58
Table 2. Analysis of the Recoveries of Hg, Pb, and As in Well, Tap, and Lake Water (n = 3).
| metal | well
water |
tap
water |
lake
water |
||||||
|---|---|---|---|---|---|---|---|---|---|
| C0 | Cf | removal | C0 | Cf | removal | C0 | Cf | removal | |
| (mg L–1) | (mg L–1) | (%) | (mg L–1) | (mg L–1) | (%) | (mg L–1) | (mg L–1) | (%) | |
| Hg | 0.88 ± 0.03 | <LDa | 100 | 2.37 ± 0.04 | <LDa | 100 | 0.81 ± 0.08 | <LDa | 100 |
| Pb | 19.55 ± 0.93 | <LDa | 100 | 6.25 ± 0.15 | <LDa | 100 | 6.11 ± 0.08 | <LDa | 100 |
| As | 13.18 ± 1.00 | 1.18 ± 0.25 | 91.05 | 15.07 ± 0.24 | <LDa | 100 | 64.17 ± 1.23 | 49.97 ± 2.06 | 22.13 |
Limit of detection.
4. Conclusions
In this work, a highly porous thiol-functionalized MIL-100(Fe) MOF was prepared by a simple grafting method. The obtained MOF was used for the dispersive solid-phase extraction of Hg(II) ions, obtaining a high extraction capacity of Hg(II), significantly higher than that of bare MIL-100(Fe), which confirmed the key role of the thiol groups on the adsorption process. Using the prepared thiol-grafted MOF, a functional MIL-100(Fe)-SH/device was obtained by an easy and straightforward coating method with a 3D printed support. The developed device, which simultaneously acts as stirrer and sorbent, could be reused efficiently over five adsorption cycles and showed good performance for the solid-phase extraction of Hg(II), Pb(II), and As(V) in real water samples, achieving recoveries of 100% for two of the ions in the three analyzed samples and demonstrating its excellent features to analyze real water samples.
Acknowledgments
The authors gratefully acknowledge the financial support from the Comunitat Autonoma de les Illes Balears through the Direcció General de Política Universitaria i Recerca with funds from the Tourist Stay Tax Law (PRD2018/45). D.R. Sáenz-García thanks the National Council of Science and Technology in Mexico, CIMAV, Santander-UIB for the scholarship.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c01544.
Operation parameters of ICP-OES and HG/CV-AFS; EDS analysis of MIL-100(Fe)-SH; FTIR spectra, SEM images, and TGA curves of MIL-100(Fe) and MIL-100(Fe)-SH; pH changes and Fe release after extraction using the MIL-100(Fe)-SH material; and zeta potential and XRD after extraction of MIL-100(Fe)-SH MOF (PDF)
Author Contributions
D.R. Sáenz-García: Investigation, Writing – original draft. Andreu Figuerola: Methodology. Gemma Turnes Palomino: Writing – review and editing, Supervision, Funding acquisition. Luz O. Leal: Writing – review and editing, Supervision. Carlos Palomino Cabello: Writing – review and editing, Visualization, Supervision.
The authors declare no competing financial interest.
Supplementary Material
References
- Saravanan A.; Senthil Kumar P.; Jeevanantham S.; Karishma S.; Tajsabreen B.; Yaashikaa P. R.; Reshma B. Effective Water/wastewater Treatment Methodologies for Toxic Pollutants Removal: Processes and Applications towards Sustainable Development. Chemosphere 2021, 280, 130595 10.1016/j.chemosphere.2021.130595. [DOI] [PubMed] [Google Scholar]
- Sall M. L.; Diaw A. K. D.; Gningue-Sall D.; Efremova Aaron S.; Aaron J. J. Toxic Heavy Metals: Impact on the Environment and Human Health, and Treatment with Conducting Organic Polymers, a Review. Environ. Sci. Pollut. Res. 2020, 27, 29927–29942. 10.1007/s11356-020-09354-3. [DOI] [PubMed] [Google Scholar]
- Jaishankar M.; Tseten T.; Anbalagan N.; Mathew B. B.; Beeregowda K. N. Toxicity, Mechanism and Health Effects of Some Heavy Metals. Interdiscipl. Toxicol. 2014, 7, 60–72. 10.2478/intox-2014-0009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kinuthia G. K.; Ngure V.; Beti D.; Lugalia R.; Wangila A.; Kamau L. Levels of Heavy Metals in Wastewater and Soil Samples from Open Drainage Channels in Nairobi, Kenya: Community Health Implication. Sci. Rep. 2020, 10, 8434. 10.1038/s41598-020-65359-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duan C.; Ma T.; Wang J.; Zhou Y. Removal of Heavy Metals from Aqueous Solution Using Carbon-based Adsorbents: a Review. J. Water Process Eng. 2020, 37, 101339 10.1016/j.jwpe.2020.101339. [DOI] [Google Scholar]
- Arikibe J. E.; Prasad S. E. Determination and Comparison of Selected Heavy Metal Concentrations in Seawater and Sediment Samples in the Coastal Area of Suva. Fiji. Mar Pollut. Bull. 2020, 157, 111157 10.1016/j.marpolbul.2020.111157. [DOI] [PubMed] [Google Scholar]
- Chen Q.; Yao Y.; Li X.; Lu J.; Zhou J.; Huang Z. Comparison of Heavy Metal Removals from Aqueous Solutions by Chemical Precipitation and Characteristics of Precipitates. J. Water Process. Eng. 2018, 26, 289–300. 10.1016/j.jwpe.2018.11.003. [DOI] [Google Scholar]
- Li J.; Hashimoto Y.; Riya S.; Terada A.; Hou H.; Shibagaki Y.; Hosomi M. Removal and Immobilization of Heavy Metals in Contaminated Soils by Chlorination and Thermal Treatment on an Industrial-scale. Chem. Eng. J. 2019, 359, 385–392. 10.1016/j.cej.2018.11.158. [DOI] [Google Scholar]
- Bashir A.; Malik L. A.; Ahad S.; Manzoor T.; Bhat M. A.; Dar G.; Pandith A. H. Removal of Heavy Metal Ions from Aqueous System by Ion-exchange and Biosorption Methods. Environ. Chem. Lett. 2019, 17, 729–754. 10.1007/s10311-018-00828-y. [DOI] [Google Scholar]
- Xiang H.; Min X.; Tang C. J.; Sillanpää M.; Zhao F. Recent Advances in Membrane Filtration for Heavy Metal Removal from Wastewater: A Mini Review. J. Water Process. Eng. 2022, 49, 103023 10.1016/j.jwpe.2022.103023. [DOI] [Google Scholar]
- Hong M.; Yu L.; Wang Y.; Zhang J.; Chen Z.; Dong L.; Zan Q.; Li R. Heavy Metal Adsorption with Zeolites: The Role of Hierarchical Pore Architecture. Chem. Eng. J. 2019, 359, 363–372. 10.1016/j.cej.2018.11.087. [DOI] [Google Scholar]
- Joseph L.; Jun B. M.; Flora J. R. V.; Park C. M.; Yoon Y. Removal of Heavy Metals from Water Sources in the Developing World Using Low-cost Materials: a Review. Chemosphere 2019, 229, 142–159. 10.1016/j.chemosphere.2019.04.198. [DOI] [PubMed] [Google Scholar]
- Huang L.; Huang W.; Shen R.; Shuai Q. Chitosan/thiol Functionalized Metal–Organic Framework Composite for the Simultaneous Determination of Lead and Cadmium Ions in Food Samples. Food Chem. 2020, 330, 127212 10.1016/j.foodchem.2020.127212. [DOI] [PubMed] [Google Scholar]
- Taghavi R.; Rostamnia S.; Farajzadeh M.; Karimi-Maleh H.; Wang J.; Kim D.; Jang H. W.; Luque R.; Varma R. S.; Shokouhimehr M. Magnetite Metal-Organic Frameworks: Applications in Environmental Remediation of Heavy Metals, Organic Contaminants, and Other Pollutants. Inorg. Chem. 2022, 61, 15747–15783. 10.1021/acs.inorgchem.2c01939. [DOI] [PubMed] [Google Scholar]
- 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 10.1016/j.ccr.2020.213554. [DOI] [Google Scholar]
- Evangelou D.; Pournara A.; Tziasiou C.; Andreou E.; Armatas G. S.; Manos M. J. Robust Al3+ MOF with Selective As(V) Sorption and Efficient Luminescence Sensing Properties toward Cr(VI). Inorg. Chem. 2022, 61, 2017–2030. 10.1021/acs.inorgchem.1c03199. [DOI] [PubMed] [Google Scholar]
- Zhang H.; Hu X.; Li T.; Zhang Y.; Xu H.; Sun Y.; Gu X.; Gu C.; Luo J.; Gao B. MIL Series of Metal Organic Frameworks (MOFs) as Novel Adsorbents for Heavy Metals in Water: A Review. J. Hazard. Mater. 2022, 429, 128271 10.1016/j.jhazmat.2022.128271. [DOI] [PubMed] [Google Scholar]
- Kobielska P. A.; Howarth A. J.; Farha O. K.; Nayak S. Metal–Organic Frameworks for Heavy Metal Removal from Water. Coord. Chem. Rev. 2018, 358, 92–107. 10.1016/j.ccr.2017.12.010. [DOI] [Google Scholar]
- Luo X.; Ding L.; Luo J. Adsorptive Removal of Pb (II) Ions from Aqueous Samples with Amino-functionalization of Metal–Organic Rrameworks MIL-101 (Cr). J. Chem. Eng. Data 2015, 60, 1732–1743. 10.1021/je501115m. [DOI] [Google Scholar]
- Wang Y.; Ye G.; Chen H.; Hu X.; Niu Z.; Ma S. Functionalized Metal–Organic Framework as a New Platform for Efficient and Selective Removal of Cadmium(II) from Aqueous Solution. J. Mater. Chem. A 2015, 3, 15292–15298. 10.1039/C5TA03201F. [DOI] [Google Scholar]
- Zhang L.; Wang J.; Du T.; Zhang W.; Zhu W.; Yang C.; Yue T.; Sun J.; Li T.; Wang J. NH2-MIL-53(Al) Metal-Organic Framework as the Smart Platform for Simultaneous High-performance Detection and Removal of Hg2+. Inorg. Chem. 2019, 58, 12573–12581. 10.1021/acs.inorgchem.9b01242. [DOI] [PubMed] [Google Scholar]
- Hu S. Z.; Huang T.; Zhang N.; Lei Y. L.; Wang Y. Enhanced Removal of Lead Ions and Methyl Orange from Wastewater Using Polyethyleneimine Grafted UiO-66-NH2 Nanoparticles. Sep. Purif. Technol. 2022, 297, 121470 10.1016/j.seppur.2022.121470. [DOI] [Google Scholar]
- Ke F.; Qiu L. G.; Yuan Y. P.; Peng F. M.; Jiang X.; Xie A. J.; Shen Y. H.; Zhu J. F. Thiol-functionalization of Metal-Organic Framework by a Facile Coordination-based Postsynthetic Strategy and Enhanced Removal of Hg2+ from Water. J. Hazard. Mater. 2011, 196, 36–43. 10.1016/j.jhazmat.2011.08.069. [DOI] [PubMed] [Google Scholar]
- Yu S.; Pang H.; Huang S.; Tang H.; Wang S.; Qiu M.; Chen Z.; Yang H.; Song G.; Fu D.; Hu B.; Wang X. Recent Advances in Metal-Organic Framework Membranes for Water Treatment: A Review. Sci. Total Environ. 2021, 800, 149662 10.1016/j.scitotenv.2021.149662. [DOI] [PubMed] [Google Scholar]
- Parsazadeh N.; Yousefi F.; Ghaedi M.; Dashtian K.; Borousan F. Preparation and Characterization of Monoliths HKUST-1 MOF via Straightforward Conversion of Cu(OH)2-based Monoliths and its Application for Wastewater Treatment: Artificial Neural Network and Central Composite Design Modeling. New J. Chem. 2018, 42, 10327–10336. 10.1039/C8NJ01067F. [DOI] [Google Scholar]
- Xu Z.; Wang J.; Li H.; Wang Y. Coating Sponge with Multifunctional and Porous Metal-Organic Framework for Oil Spill Remediation. Chem. Eng. J. 2019, 370, 1181–1187. 10.1016/j.cej.2019.03.288. [DOI] [Google Scholar]
- Zhao R.; Shi X.; Ma T.; Rong H.; Wang Z.; Cui F.; Zhu G.; Wang C. Constructing Mesoporous Adsorption Channels and MOF-polymer Interfaces in Electrospun Composite Fibers for Effective Removal of Emerging Organic Contaminants. ACS Appl. Mater. Interfaces 2021, 13, 755–764. 10.1021/acsami.0c20404. [DOI] [PubMed] [Google Scholar]
- Zhang D.; Xiao J.; Guo Q.; Yang J. 3D-Printed Highly Porous and Reusable Chitosan Monoliths for Cu(II) Removal. J. Mater. Sci. 2019, 54, 6728–6741. 10.1007/s10853-019-03332-y. [DOI] [Google Scholar]
- Pianca D.; Carboni M.; Meyer D. 3D-Printing of Porous Materials: Application to Metal-Organic Frameworks. Mater. Lett. X 2022, 13, 100121 10.1016/j.mlblux.2022.100121. [DOI] [Google Scholar]
- Hartings M. R.; Ahmed Z. Chemistry from 3D Printed Objects. Nat. Rev. Chem. 2019, 3, 305–314. 10.1038/s41570-019-0097-z. [DOI] [Google Scholar]
- Waheed S.; Rodas M.; Kaur H.; Kilah N. L.; Paull B.; Maya F. In-situ Growth of Metal-Organic Frameworks in a Reactive 3D Printable Material. Appl. Mater. Today 2021, 22, 100930 10.1016/j.apmt.2020.100930. [DOI] [Google Scholar]
- Kearns E. R.; Gillespie R.; D’Alessandro D. M. 3D Printing of Metal–Organic Framework Composite Materials for Clean Energy and Environmental Applications. J. Mater. Chem. A 2021, 9, 27252–27270. 10.1039/D1TA08777K. [DOI] [Google Scholar]
- Mallakpour S.; Azadi E.; Hussain C. M. 26 MOF/COF-based Materials Using 3D Printing Technology: Applications in Water Treatment, Gas Removal, Biomedical, and Electronic Industries. New J. Chem. 2021, 45, 13247–13257. 10.1039/D1NJ02152D. [DOI] [Google Scholar]
- Li R.; Yuan S.; Zhang W.; Zheng H.; Zhu W.; Li B.; Zhou M.; Wing-Keung Law A.; Zhou K. 3D Printing of Mixed Matrix Films Based on Metal-Organic Frameworks and Thermoplastic Polyamide 12 by Selective Laser Sintering for Water Applications. ACS Appl. Mater. Interfaces 2019, 11, 40564–40574. 10.1021/acsami.9b11840. [DOI] [PubMed] [Google Scholar]
- Horcajada P.; Surblé S.; Serre C.; Hong D. Y.; Seo Y. K.; Chang J. S.; Grenèche J. M.; Margiolaki I.; Férey G. Synthesis and Catalytic Properties of MIL-100(Fe), an Iron(III) Carboxylate with Large Pores. Chem. Commun. 2007, 2820–2822. 10.1039/B704325B. [DOI] [PubMed] [Google Scholar]
- Yuan B.; Wang X.; Zhou X.; Xiao J.; Li Z. Novel Room-temperature Synthesis of MIL-100(Fe) and its Excellent Adsorption Performances for Separation of Light Hydrocarbons. Chem. Eng. J. 2019, 355, 679–686. 10.1016/j.cej.2018.08.201. [DOI] [Google Scholar]
- Denny M. S.; Cohen S. M. In Situ Modification of Metal-Organic Frameworks in Mixed-matrix Membranes. Angew. Chem., Int. Ed. 2015, 54, 9029–9032. 10.1002/anie.201504077. [DOI] [PubMed] [Google Scholar]
- Nehra M.; Dilbaghi N.; Singhal N. K.; Hassan A. A.; Kim K. H.; Kumar S. Metal Organic Frameworks MIL-100(Fe) as an efficient adsorptive material for phosphate management. Environ. Res. 2019, 169, 229–236. 10.1016/j.envres.2018.11.013. [DOI] [PubMed] [Google Scholar]
- Ryu J.; Lee M. Y.; Song M. G.; Baeck S. H.; Shim S. E.; Qian Y. Highly Selective Removal of Hg(II) Ions from Aqueous Solution Using Thiol-modified Porous Polyaminal-networked Polymer. Sep. Purif. Technol. 2020, 250, 117120 10.1016/j.seppur.2020.117120. [DOI] [Google Scholar]
- Ji C.; Ren Y.; Yu H.; Hua M.; Lv Lu.; Zhang W. Highly Efficient and Selective Hg(II) Removal from Water by Thiol-functionalized MOF-808: Kinetic and Mechanism Study. Chem. Eng. J. 2022, 430, 132960 10.1016/j.cej.2021.132960. [DOI] [Google Scholar]
- Leclerc H.; Vimont A.; Lavalley J. C.; Daturi M.; Wiersum A. D.; Llwellyn P. L.; Horcajada P.; Férey G.; Serre C. Infrared Study of the Influence of Reducible Iron(III) Metal Sites on the Adsorption of CO, CO2, Propane, Propene and Propyne in the Mesoporous Metal–Organic Framework MIL-100. Phys. Chem. Chem. Phys. 2011, 13, 11748–11756. 10.1039/c1cp20502a. [DOI] [PubMed] [Google Scholar]
- Yoon J. W.; Seo Y. K.; Hwang Y. K.; Chang J. S.; Leclerc H.; Wuttke S.; Bazin P.; Vimont A.; Daturi M.; Bloch E.; Llewellyn P. L.; Serre C.; Horcajada P.; Grenèche J. M.; Rodrigues A. E.; Férey G. Controlled Reducibility of a Metal–Organic Framework with Coordinatively Unsaturated Sites for Preferential Gas Sorption. Angew. Chem., Int. Ed. 2010, 49, 5949–5952. 10.1002/anie.201001230. [DOI] [PubMed] [Google Scholar]
- Horcajada P.; Chevreau H.; Heurtaux D.; Benyettou F.; Salles F.; Devic T.; Garcia-Marquez A.; Yu C.; Lavrard H.; Dutson C. L.; Magnier E.; Maurin G.; Elkaïm E.; Serre C. Extended and Functionalized Porous Iron(III) tri- or Dicarboxylates with MIL-100/101 Topologies. Chem. Comm. 2014, 50, 6872–6874. 10.1039/c4cc02175d. [DOI] [PubMed] [Google Scholar]
- Zhong R.; Yu X.; Meng W.; Liu J.; Zhi C.; Zou R. Amine-grafted MIL-101(Cr) via Double-solvent Incorporation for Synergistic Enhancement of CO2 Uptake and Selectivity. ACS Sustain. Chem. Eng. 2018, 6, 16493–16502. 10.1021/acssuschemeng.8b03597. [DOI] [Google Scholar]
- Pirzadeh K.; Ghoreyshi A. A.; Rohani S.; Rahimnejad M. Strong Influence of Amine Grafting on MIL-101 (Cr) Metal-Organic Framework with Exceptional CO2/N2 Selectivity. Ind. Eng. Chem. Res. 2020, 59, 366–378. 10.1021/acs.iecr.9b05779. [DOI] [Google Scholar]
- Yantasee W.; Warner C. L.; Sangvanich T.; Addleman R. S.; Carter T. G.; Wiacek R. J.; Fryxell G. E.; Timchalk C.; Warner M. G. Removal of Heavy Metals from Aqueous Systems with Thiol Functionalized Superparamagnetic Nanoparticles. Environ. Sci. Technol. 2007, 41, 5114–5119. 10.1021/es0705238. [DOI] [PubMed] [Google Scholar]
- Figueira P.; Lourenço M. A. O.; Pereira E.; Gomes J. R. B.; Ferreira P.; Lopes C. B. Periodic Mesoporous Organosilica with Low Thiol Density – a Safer Material to Trap Hg(II) from Water. J. Environ. Chem. Eng. 2017, 5, 5043–5053. 10.1016/j.jece.2017.09.032. [DOI] [Google Scholar]
- Li Y.; Tan M.; Liu G.; Si D.; Chen N.; Zhou D. Thiol-functionalized Metal–Organic Frameworks Embedded with Chelator-modified Magnetite for High-efficiency and Recyclable Mercury Removal in Aqueous Solutions. J. Mater. Chem. A 2022, 10, 6724–6730. 10.1039/D1TA10906E. [DOI] [Google Scholar]
- Liang R.; Zou H. Removal of Aqueous Hg(II) by Thiol-functionalized Nonporous Silica Microspheres Prepared by One-step Sol–gel Method. RSC Adv. 2020, 10, 18534–18542. 10.1039/D0RA02759F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Merí-Bofí L.; Royuela S.; Zamora F.; Ruiz-González M. L.; Segura J. L.; Muñoz-Olivas R.; Mancheño M. J. Thiol Grafted Imine-based Covalent Organic Frameworks for Water Remediation Through Selective Removal of Hg(II). J. Mater. Chem. A 2017, 5, 17973–17981. 10.1039/C7TA05588A. [DOI] [Google Scholar]
- Sohrabi M. R. Preconcentration of Mercury(II) Using a Thiol-functionalized Metal-Organic Framework Nanocomposite as a Sorbent. Microchim. Acta 2014, 181, 435–444. 10.1007/s00604-013-1133-1. [DOI] [Google Scholar]
- Bezverkhyy I.; Weber G.; Bellat J. P. Degradation of Fluoride-free MIL-100(Fe) and MIL-53(Fe) in Water: Effect of Temperature and pH. Microporous Mesoporous Mater. 2016, 219, 117–124. 10.1016/j.micromeso.2015.07.037. [DOI] [Google Scholar]
- Zhang W.; An Y.; Li S.; Liu Z.; Chen Z.; Ren Y.; Wang S.; Zhang X.; Wang X. Enhanced Heavy Metal Removal from an Aqueous Environment Using an Eco-friendly and Sustainable Adsorbent. Sci. Rep. 2020, 10, 1–19. 10.1038/s41598-020-73570-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gomaa H.; Shenashen M. A.; Yamaguchi H.; Alamoudi A. S.; Abdelmottaleb M.; Cheira M. F.; Seaf El-Naser T. A.; El-Safty S. A. Highly-Efficient Removal of AsV, Pb2+, Fe3+, and Al3+ Pollutants from Water Using Hierarchical, Microscopic TiO2 and TiOF2 Adsorbents through Batch and Fixed-Bed Columnar Techniques. J. Cleaner Prod. 2018, 182, 910–925. 10.1016/j.jclepro.2018.02.063. [DOI] [Google Scholar]
- Yap P. L.; Tung T. T.; Kabiri S.; Matulick N.; Tran D. N. H.; Losic D. Polyamine-Modified Reduced Graphene Oxide: A New and Cost-Effective Adsorbent for Efficient Removal of Mercury in Water. Sep. Purif. Technol. 2020, 238, 116441 10.1016/j.seppur.2019.116441. [DOI] [Google Scholar]
- Ragheb E.; Shamsipur M.; Jalali F.; Mousavi F. Modified Magnetic-Metal Organic Framework as a Green and Efficient Adsorbent for Removal of Heavy Metals. J. Environ. Chem. Eng. 2022, 10, 107297 10.1016/j.jece.2022.107297. [DOI] [Google Scholar]
- Kumar A.; Das T.; Thakur R. S.; Fatima Z.; Prasad S.; Ansari N. G.; Patel D. K. Synthesis of Biomass-Derived Activated Carbons and Their Immobilization on Alginate Gels for the Simultaneous Removal of Cr(VI), Cd(II), Pb(II), As(III), and Hg(II) from water. ACS Omega 2022, 7, 41997–42011. 10.1021/acsomega.2c03786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saldaña-Robles A.; Damian-Ascencio C. E.; Guerra-Sanchez R. J.; Saldaña-Robles A. L.; Saldaña-Robles N.; Gallegos-Muñoz A.; Cano-Andrade S. Effects of the Presence of Organic Matter on the Removal of Arsenic from Groundwater. J. Clean Prod. 2018, 183, 720–728. 10.1016/j.jclepro.2018.02.161. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.


