Ag-exchanged mesoporous chromium terephthalate with sulfonate for removing radioactive methyl iodide at extremely low concentrations in humid environments

https://doi.org/10.1016/j.jhazmat.2021.125904Get rights and content

Highlights

  • Ag-exchanged mesoporous MIL-101 materials with a sulfonate group are prepared for the effective removal of methyl iodide.

  • Enhanced hydrophobicity of Ag-MIL-101 functionalized by alkyl sulfonate group is confirmed by the in situ FT-IR and HI.

  • Outstanding performance for CH3I removal is achieved even under RH 95% and at the ppb CH3I concentration level.

Abstract

The development of efficient adsorbents to remove radioactive methyl iodide (CH3I) in humid environments is crucial for air purification after pollution by nuclear power plant waste. In this work, we successfully prepared a post-synthetic covalent modified MIL-101 with a sulfonate group followed by the ion-exchange of Ag (I), which is well characterized by diffuse reflectance FT-IR spectroscopy, X-ray photoelectron spectroscopy (XPS) and the hydrophobic index (HI). After modification of the MOFs, we applied functionalized MIL-101 obtained by either one-pot synthesis (MIL-101-SO3Ag) or a post-synthetic modification process (MIL-101-RSO3Ag, R = NH(CH2)3) to remove the CH3I at an extremely low concentration (0.31 ppm) in an environment with very high relative humidity (RH 95%). Enhanced hydrophobicity of the surface-modified MIL-101 was evaluated by examining the HI with the competitive adsorption of water and cyclohexane vapor, with a high surface area maintained, as confirmed by Ar physisorption. Interestingly, the post-synthetically modified MIL-101-RSO3Ag showed exceptional adsorption performance as determined by its decontamination factor (DF = 195,350) at 303 K and RH 95%. This performance was in comparison to Ag (I)-exchanged 13X zeolite and MIL-101-SO3Ag, which include much higher amounts of Ag. Furthermore, MIL-101-RSO3Ag retained ~94–100% of its fresh adsorbent performance during five cycle repetitions.

Introduction

The effective management of several fission products released from nuclear power plants is one of the major issues related to the safe use of nuclear energy (Nandanwar et al., 2016). Among radionuclides, isotopes of radioactive iodine (129I and 131I) have a critical impact on the human body and on the environment due to nuclear fission properties, especially considering the high radioactivity of 131I with a short half-life of 8.05 days and the accumulation and persistence of 129I in the environment given its long half-life of 1.52 × 107 years (Nandanwar et al., 2016, Dickman et al., 2003). Gaseous radionuclides in elemental (I2) and organic iodine (CH3I) forms can be released not only under accidental conditions but also even during normal power operations (Nandanwar et al., 2016, Nenoff et al., 2014, Reyerson and Cameron, 1936, Choi et al.,). Specifically, gaseous radioactive iodine released under normal power operations is of great concern owing to the low concentration of radionuclides as well as the typically high humid of these environments (Nenoff et al., 2014, Choi et al.,). Elemental iodine (I2) can easily be removed by conventional adsorbents such as activated carbon and silver-doped materials, even under extremely humid conditions. However, it is known that the removal of organic iodides by the adsorbents is more difficult than for the elemental iodine and that the removal efficiency can be steeply decreased in humid environments (Tietze et al., 2013, Hughes et al., 2013). As a results, an enormous research effort has been invested to develop novel adsorbent materials that can efficiently remove organic iodides, methyl iodide (CH3I) in particular, in humid environments.

ASZM-TEDA (an activated carbon impregnated with Ag, Cu, Zn, and Mo), triethylenediamine (TEDA) (Park et al., 1995, Aneheim et al., 2018, Park et al., 2001), and silver-exchanged porous materials such as silica gel (Asmussen et al., 2019) or zeolites (Nenoff et al., 2014, Pham et al., 2016, Chebbi et al., 2017, Azambre and Chebbi, 2017, Chebbi et al., 2016, Chapman et al., 2010) are representative adsorbents used for CH3I removal. However, these materials exhibited limitations when used for competitive adsorption at a relative humidity (RH) of 95% at 303 K and extremely low concentration (0.31 ppm) of CH3I. The commercial adsorbent ASZM-TEDA provided sufficient amine adsorption sites (in TEDA), but the force binding ASZM and TEDA was not high enough for effective competitive adsorption of CH3I in humid environments. In this experimental condition, the molar ratio of methyl iodide to water is 1: 130,000. Accordingly, the presence of active sites that can more selectively bind to CH3I becomes crucial. Alternatively, porous materials with Ag (I) incorporated, such as zeolites capable of strong alkylation leading to the formation of AgI, have been applied to increase the adsorption performance of CH3I removal. Although Ag-exchanged zeolite was applied to remove CH3I, limitations such as the high loading amount of Ag (~23 wt%) and lower adsorption capacity at RH 95% are still unresolved (Chebbi et al., 2017, Azambre and Chebbi, 2017, Chebbi et al., 2016).

Metal-organic frameworks (MOFs) with framework structures comprising metal- containing nodes connected by organic bridging ligands, have proven to offer advantages as promising adsorbents. This is due to their high surface area, exceptional porosity, tunable pore size, and simple modulation. In general, MOFs has been extensively studied due to the diversity of their framework designs, which provide a great variety of pore structures and properties using the post-synthetic modification (PSM) approach (Cohen, 2012). MOFs are receiving close attention in the fields of gas sorption (Rosi et al., 2003, Yoon et al., 2017, Kaye et al., 2007), gas separation (Couck et al., 2009, Liu et al., 2018, Li et al., 2009), sensing (Kreno et al., 2012), and catalysis (Lee et al., 2009, Farrusseng et al., 2009, Valerkar et al., 2016). The PSM process for the synthesis of MOF has been regarded as a very promising method for the introduction of functionality in MOF frameworks without causing structural damage to the parent MOF. Moreover, post-synthetically modified MOFs have been used as adsorbents in the removal of toxic industrial gases, chemical warfare agents, and radioactive gases (Sava et al., 2011, Zeng et al., 2010, Banerjee et al., 2016, Sava et al., 2013, Li et al., 2017). Recently, MOFs including square forms of HKUST-1 and MIL-53, and the hexagonal window geometry of MIL-120, have been reported to exhibit the best CH3I adsorption capacities (425, 164, and 127 mg/g) at 308 K and 1333 ppm of CH3I (Chebbi et al., 2018).

Very recently, Li et al. reported the post-synthetic functionalization of MIL-101 with different molecules having amine groups. Among the various amine-functionalized forms of MIL-101, MIL-101-TEDA showed a higher loading capacity (120 wt%) of CH3I within 10 min and reached its maximum uptake amount of 160 wt% by 120 min at 313 K (Li et al., 2017). The authors have also evaluated the efficiency of adsorbents with the “decontamination factor” (DF), as the ratio of radioactivity before and after decontamination procedures. For a low CH3I concentration of 50 ppm, the DF of MIL-101-TEDA has been achieved within the range 4800–6300 at 423 K. However, the adsorption data of MIL-101-TEDA has not been determined at an extremely low concentration (0.31 ppm) of CH3I, which is a critical point related to the off-gas management of the facilities of a nuclear power plant under normal operating conditions.

In the work reported herein, we prepared a series of Brønsted acid-functionalized MIL-101s, made using either one-pot synthesis (Akiyama et al., 2011) or the sequential PSM method (Andriamitantsoa et al., 2016). By introducing sulfonic acid groups through two strategies, they could be utilized for the ion exchange of Ag, which can strongly interact with CH3I. MIL-101-SO3H was synthesized in a one-pot fashion using a ligand with a sulfonate group, while MIL-101-RSO3H (R = NH(CH2)3) was prepared via consecutive PSM processes (MIL-101-NO2 to -NH2, to -NH-RSO3H) incorporating more hydrophobic alkyl sulfonate to minimize the effect of humidity. Both samples were further modified using the Ag ion-exchange process, and the CH3I removal efficiency of MIL-101-RSO3Ag was markedly improved (by more than 99.99%) at 303 K, even at RH 95%. Furthermore, we demonstrated that the Ag ion-exchanged MIL-101-RSO3H could be a promising candidate for use under the real conditions in a nuclear power plant.

Section snippets

Chemicals

Terephthalic acid (TPA, 98.0%), 2-nitroterephthalic acid (TPA-NO2, 99.0%), CrCl3.6H2O (98.0%), Cr(NO3)3.9H2O (99.0%), SnCl2 (98.0%), AgNO3 (99.0%), 1,4-diazabicyclo[2.2.2]octane (TEDA, 99.0%), hydrofluoric acid (HF, 48.0%), cyclohexane (99.5%), thiophene (99.0%) and 13X zeolite were obtained from Sigma-Aldrich. Hydrochloric acid (HCl, 37.0%), acetonitrile (CH3CN, 99.5%), chloroform (CHCl3, 99.9%), and ethanol (95.0%) were purchased from Samchun Pure Chemicals (South Korea) and used without

Post synthetic modification of MIL-101

MIL-101 with the structural formula Cr33-O)F(H2O)2(O2CC6H4CO2)3 has outstanding capability for adsorption of molecules due to its high thermal and chemical stability and large apparent surface area (Férey et al., 2005). The framework structure of MIL-101 contains two different types of large mesoporous cages with free diameters of approximately 2.9 and 3.4 nm, accessible through pentagonal (1.2 nm) and hexagonal (1.4 nm) windows, respectively. These textural features, along with its high

Conclusions

In this work, two strategies were implemented to improve the removal efficiencies of methyl iodide under humid conditions. This was accomplished through the exchange of an Ag ion for a sulfonate group on surface-modified MOFs. For use of this technology in normal power operations, it is important to develop an adsorbent with a higher removal rate (99.99% or more) due to the property of bioaccumulation shown by radioactive iodine in the body. At low concentrations, chemisorption is more

CRediT authorship contribution statement

Ga-Young Cha: Methology, Investigation, Writing-Original draft, Validation. E. S. Sanil: Methology, Writing-Original draft. Mijung Lee: Methology, Formal analysis. Kyung-Ryul Oh: Methology, Formal analysis. Anil H. Valekar: Methology. Min-Kun Kim: Methology, Formal analysis. Heesoo Jung: Methology, Formal analysis. Do-Young Hong: Methology, Conceptualization, Writing-Review and Editing. Young Kyu Hwang: Conceptualization, Writing-Original draft, Writing-Review and Editing, Funding acquisition.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reports in this paper.

Acknowledgements

This work was supported by the Agency for Defense Development of Korea (UC190034GD), Republic of Korea and the Defense Industry Technology Center (DITC) of Korea (UC5000ID), Republic of Korea. YKH are kindly grateful to the Institutional Research Program of KRICT (SI-1911-20), Republic of Korea for financial support.

References (52)

  • L. Wu et al.

    Effect of liquid-phase O3 oxidation of activated carbon on the adsorption of thiophene

    Chem. Eng. J.

    (2014)
  • G. Akiyama et al.

    Cellulose hydrolysis by a new porous coordination polymer decorated with sulfonic acid functional groups

    Adv. Mater.

    (2011)
  • R.S. Andriamitantsoa et al.

    SO3H-functionalized metal organic frameworks: an efficient heterogeneous catalyst for the synthesis of quinoxaline and derivatives

    RSC Adv.

    (2016)
  • B. Azambre et al.

    Evaluation of silver zeolites sorbents toward their ability to promote stable CH3I storage as AgI precipitates

    ACS Appl. Mater. Interfaces

    (2017)
  • D. Banerjee et al.

    Metal–organic framework with optimally selective xenon adsorption and separation

    Nat. Commun.

    (2016)
  • S. Ban et al.

    Thiophene separation with silver-doped Cu-BTC metal-organic framework for deep desulfurization

    Ind. Eng. Chem. Res.

    (2018)
  • S. Bernt et al.

    Direct covalent post-synthetic chemical modification of Cr-MIL-101 using nitrating acid

    Chem. Comm.

    (2011)
  • B.N. Bhadra et al.

    Liquid-phase adsorption of aromatics over a metal–organic framework and activated carbon: effects of hydrophobicity/hydrophilicity of adsorbents and solvent polarity

    J. Phys. Chem. C.

    (2015)
  • K.W. Chapman et al.

    Radioactive iodine capture in silver-containing mordenites through nanoscale silver iodide formation

    J. Am. Chem. Soc.

    (2010)
  • M. Chebbi et al.

    A Combined DRIFTS and DR-UV–Vis spectroscopic in situ study on the trapping of CH3I by silver-exchanged faujasite zeolite

    J. Phys. Chem. C

    (2016)
  • B.S. Choi, S.-B. Kim, J. Moon, B.-K. Seo, Evaluation of decontamination factor of radioactive methyl iodide on...
  • H. Chun et al.

    Universal scaling relationship to screen an efficient metallic adsorbent for adsorptive removal of iodine gas under humid conditions: first-principles study

    J. Phys. Chem. C.

    (2018)
  • S.M. Cohen

    Postsynthetic methods for the functionalization of metal–organic frameworks

    Chem. Rev.

    (2012)
  • S. Couck et al.

    An amine-functionalized MIL-53 metal-organic framework with large separation power for CO2 and CH4

    J. Am. Chem. Soc.

    (2009)
  • S. Devautour‐Vinot et al.

    Guest‐assisted proton conduction in the sulfonic mesoporous MIL‐101 MOF

    Chem.– Asian J.

    (2019)
  • P.W. Dickman et al.

    Thyroid cancer risk after thyroid examination with I-131: a population-based cohort study in Sweden

    Int. J. Cancer

    (2003)
  • Cited by (17)

    • Adsorptive desulfurization using Cu<sup>+</sup>modified UiO-66(Zr) via ethanol vapor reduction

      2022, Journal of Environmental Chemical Engineering
      Citation Excerpt :

      But the crystal size of UiO-66-SO3Cu(I)-X began to increase with increasing the molar ratio of sulfonic acid groups to total organic ligand. After the increase in the number of sulfonic acid groups added during the synthesis process, it leads to more exchangeable cation positions in the skeleton, which affects the size and morphology of the particles [39,40]. So, we can see the phenomenon of rough surface of the material in the Fig. 4.

    • CO<inf>2</inf> capture performance of fluorinated porous carbon composite derived from a zinc-perfluoro metal-organic framework

      2022, Separation and Purification Technology
      Citation Excerpt :

      Competitive adsorption of water as hydrophilic moiety and cyclohexane as hydrophobic moiety was carried out. Detailed experimental [37–39] was added in supporting information. Zn-perfluoro derivatives, and activated carbon were shaped into sphere type with 1 wt% of polyvinyl butyral in ethanol solvent.

    View all citing articles on Scopus
    View full text