U(VI) adsorption to Fe3O4 nanoparticles coated with lignite humic acid: Experimental measurements and surface complexation modeling

https://doi.org/10.1016/j.colsurfa.2021.126150Get rights and content

Highlights

  • U(VI) adsorption onto LHA-coated Fe3O4 NPs increases with pH and LHA content.

  • Surface complexation model developed to describe and predict U(VI) adsorption.

  • Several binding sites and aqueous uranyl species involved in U(VI) adsorption.

  • U(VI) binds almost exclusively onto LHA-hosted carboxyl binding sites.

  • LHA-coated Fe3O4 NPs are promising sorbents for U(VI) removal from solutions.

Abstract

Lignite humic acid-coated Fe3O4 nanoparticles (LHA-coated Fe3O4 NPs) with different extents of LHA coating were synthesized and characterized by scanning electron microscopy (SEM), X-ray diffraction (XRD), Fourier transformed infrared (FTIR), X-ray photoelectron spectroscopy (XPS) and potentiometric titration. The results of characterization revealed that LHA-coated Fe3O4 NPs contained plentiful oxygen-containing groups. Batch adsorption measurements show that U(VI) adsorption is pH dependent, with an optimum pH range of 5.0–8.0. The removal kinetics of U(VI) can be well fitted by the pseudo-second-order model. The maximum adsorption capacities at 298 K were 42.5, 55.6, and 68.7 mg g−1 on 0.5, 1.5, and 2.5 LHA-coated Fe3O4 NPs, respectively. The regeneration experiment showed that LHA-coated Fe3O4 NPs exhibited good recoverability for U(VI) adsorption. The adsorption mechanism of U(VI) on LHA-coated Fe3O4 NPs revealed by XPS and surface complexation model. XPS results indicate that both > Fe surface sites and oxygen-containing groups are involved in U(VI) adsorption. Moreover, no significant U(VI) reduction occurs on the surfaces of the LHA-coated Fe3O4 NPs, possibly because of the insulating properties of the LHA coating. Surface complexation modeling further demonstrated that U(VI) adsorption onto the Fe3O4 NPs with the highest extents of LHA coating is dominated by binding onto the LHA-associated binding sites (e.g., carboxyl sites), whereas > Fe surface sites are also involved in U(VI) adsorption onto Fe3O4 NPs with the lowest extent of LHA coating. LHA-coated Fe3O4 NPs represents a cost-efficient and good stability adsorbent for the effective removal of U(VI) from aqueous solutions, and our results provide the means for modeling the adsorption behavior as a function of the pH and extent of LHA coating.

Introduction

Uranium-contaminated drainage water from uranium mine tailings and waste disposal from nuclear fuel processing have affected the surface water, groundwater, and soils at a number of sites worldwide [[1], [2], [3]]. Under oxic conditions, uranium is present in a solution as a highly mobile U(VI) species [4], and poses a potential threat to the environment and human health owing to its radioactivity and biological toxicity. In 2011, the World Health Organization established a provisional guideline that uranium in drinking water should not exceed 30 μg L−1 [5], and a large number of studies over the past decades have focused on the engineered removal or enrichment of U(VI) from aqueous solutions [[6], [7], [8], [9], [10]].

The aqueous uranyl cation (UO22+) can adsorb onto iron oxides [[11], [12], [13], [14]], and microscale magnetite and magnetite nanoparticles (Fe3O4 NPs) have attracted significant attention for use in engineered adsorption processes [[15], [16], [17], [18], [19], [20]] because they can be quickly recovered through a magnetic separation. However, uncoated Fe3O4 NPs are prone to aggregation [21], which affects their efficiency and restricts their large-scale application in environmental remediation. The surface modification of Fe3O4 NPs can reduce the particle aggregation, and improve the reaction performance [[22], [23], [24]]. For example, Pan et al. [25] measured the U(VI) adsorption onto Fe3O4 NPs functionalized with three separate simple organic acids (stearic acid, oleic acid, and octadecylphosphonic acid), finding that organic acid coatings are highly effective in preventing the aggregation of Fe3O4 NPs, and that the highest adsorption densities reach 103 mg g−1 when equilibrated with a total U(VI) concentration of 9.4 μM at pH 6–7.

Pan et al. [25] demonstrated the effectiveness of organic acid coatings on Fe3O4 NPs in preventing the aggregation and enhancing the metal adsorption. However, Pan et al. [25] used simple laboratory-grade organic acids to coat their NPs, whereas natural organic matter (NOM) would be a less expensive coating material. Furthermore, NOM possesses a wide range and high concentration of functional group types, potentially enhancing the binding capability relative to single-site organic acids. Humic acid (HA) is ubiquitous in soils, surface water, and groundwater systems [26] and can be used as an inexpensive means of coating NPs to decrease the extent of NP aggregation through electrostatic repulsive forces. HA forms strong bonds with heavy metal owing to its abundant carboxylic, phenolic, and sulfhydryl functional groups; therefore, HA coatings on NPs can lead to the effective adsorption removal of heavy metals from a solution [26,27]. HA-coated Fe3O4 NPs have been previously prepared and studied and are promising materials for the removal of environmental pollutants [[28], [29], [30], [31]]. Despite there being a number of studies of metal adsorption onto HA-coated Fe3O4 NPs, few studies have examined U(VI) removal using such NPs. Singhal et al. [32] synthesized Fe3O4 NPs with different extents of HA surface coatings and measured the sorption of U(VI) onto these materials using DI water and sea water matrices, and found that an optimum extent of HA coating exists. However, the relative contributions of the Fe3O4 and HA binding sites in the overall removal of U(VI) were unconstrained in their study. In addition, previous studies have used commercially supplied HAs as NP coatings [[30], [31], [32]], without knowing the exact source or characteristics of the HA applied. The chemical properties of different NOM vary widely, and hence, the effect of NOM coatings on mineral particles on the metal adsorption behaviors likely is strongly NOM-specific [33]. Furthermore, all previous studies on metal adsorption onto HA-coated Fe3O4 NPs have used a Langmuir or Freundlich modeling approach. Because the Langmuir and Freundlich modeling parameters vary as a function of the system conditions (e.g., pH and solution composition, etc.), surface complexation modeling (SCM) represents a more flexible tool for predicting the metal adsorption under conditions not directly studied in the laboratory [[34], [35], [36], [37]]. For example, Pan et al. [25] used a diffuse double-layer SCM approach to model U(VI) adsorption onto Fe3O4 NPs coated with simple organic acids and found that a series of uranyl-, uranyl-hydroxide, and uranyl-carbonate surface complexes are required to account for the adsorption behavior observed. However, no studies have attempted to use an SCM approach to model U(VI) adsorption onto Fe3O4 NPs coated with complex HA.

In this study, we measured U(VI) adsorption onto synthesized Fe3O4 NPs coated with different amounts of HA derived from lignite (LHA). Lignite is a type of low-grade and low-value coal, and there are estimated to be more than one trillion tons of it across the world [38]. Because of the abundance of lignite, LHA is less expensive than commercially available HAs. Similar to other HAs, LHA contains predominantly carboxylic and phenolic functional groups [39,40], but its higher sorption capacity for heavy metals [41] suggests that LHA contains a higher concentration of binding sites than in HAs, despite the site concentrations on LHA having not been previously measured. Thus, LHA can bind more tightly onto the surface of the NPs when used as a coating material, thereby increasing the stability of the NP adsorbents in the suspension. Despite the potential advantages of LHA, most research and applications of HA coatings onto NPs have involved commercial HA, and relatively few studies have been conducted with LHA-coated NPs. In this study, we measured the adsorption behavior as a function of the solution pH, ionic strength, reaction time, and sorbate-to-sorbent ratio. XPS measurements of LHA-coated Fe3O4 NPs before and after U(VI) adsorption were collected to probe a possible change in the valence state of uranium accompanying the adsorption. To understand the relative contributions of the Fe3O4 NPs and the LHA components during U(VI) removal, we used a non-electrostatic SCM to interpret the adsorption data, and we conducted experimental measurements to determine the stability constants for the important surface complexes.

Section snippets

Sorbent preparation and characterization

LHA was extracted from Lincang lignite from Yunnan Province, China, and purified through a modification of the traditional alkaline-acid method, as reported previously [26]. Briefly, the lignite sample was mixed with 0.5 mol L−1 NaOH and 0.1 mol L−1 Na4P2O7·10H2O (1:1) at a volume ratio of 10:1 (m/V) and shaken for 3 h. Then, 6 mol L−1 HCl was added to the supernatant until obtaining a pH of 1 to precipitate the HAs. Next, a 0.5 % HCl-HF solution was added to the precipitated HAs to eliminate

Characterization of samples

The FT-IR spectra of the Fe3O4 NPs, LHA-coated Fe3O4 NPs, and LHA before and after adsorption are shown in Fig. 1a and b. The peak at ∼580 cm−1 in the LHA-coated Fe3O4 NP samples is related to the stretching vibration of the Fe-O bond [48]. However, the intensity of the Fe-O bond becomes weaker with an increase in the LHA concentration, which is likely because the increased LHA concentration obscures the signal of the Fe-O bond. The peaks at ∼3410 cm-1 in each of the samples are related to OHsingle bond

Conclusions

In this study, we measured the adsorption behavior of LHA-coated Fe3O4 NPs for the removal of U(VI) from a solution under a wide range of conditions. U(VI) adsorption onto LHA-coated Fe3O4 NPs was pH dependent, with an optimum pH range of 5.0–8.0. The LHA-coated Fe3O4 NPs adsorbed significantly higher concentrations of U(VI) than the uncoated Fe3O4 NPs. Specifically, the maximum adsorption capacities at pH 5.0 were 42.5, 55.6, and 68.7 mg g−1 on 0.5, 1.5, and 2.5 LHA-coated Fe3O4 NPs,

CRediT authorship contribution statement

Yangyang Zhang: Conceptualization, Investigation, Data curation, Visualization, Writing. Jeremy B. Fein: Methodology, Software, Formal analysis, Funding acquisition, Writing - review & editing, Project administration. Yilian Li: Methodology. Qiang Yu: Software, Formal analysis. Bo Zu: Project administration, Supervision, Writing - review & editing. Chunli Zheng: Supervision, Writing - review & editing.

Declaration of Competing Interest

The authors declare no conflict of interest.

Acknowledgements

The DOC analysis and some of the U analyses were conducted at the Center for Environmental Science and Technology (CEST) at the University of Notre Dame. This work was supported by a scholarship from China Scholarship Council [CSC student ID: 201806410028], a research fund from Chongqing Jiaotong University (20JDKJC-B007) and Natural Science Foundation of Chongqing, China (cstc2020jcyj-msxmX0763).

References (68)

  • A. Tripathi et al.

    Uranium (VI) recovery from aqueous medium using novel floating macroporous alginate-agarose-magnetite cryobeads

    J. Hazard. Mater.

    (2013)
  • W. Zhu et al.

    Procedural growth of fungal hyphae/Fe3O4/graphene oxide as ordered-structure composites for water purification

    Chem. Eng. J.

    (2019)
  • T. Missana et al.

    Uranium (VI) sorption on colloidal magnetite under anoxic environment: experimental study and surface complexation modelling

    Geochim. Cosmochim. Acta.

    (2003)
  • K. Zhu et al.

    Solvent-free engineering of Fe°/Fe3C nanoparticles encased in nitrogen-doped carbon nanoshell materials for highly efficient removal of uranyl ions from acidic solution

    J. Colloid Interface Sci.

    (2020)
  • D. Maity et al.

    Synthesis of iron oxide nanoparticles under oxidizing environment and their stabilization in aqueous and non-aqueous media

    J. Magn. Magn. Mater.

    (2007)
  • L. Tan et al.

    Synthesis of Fe3O4@TiO2 core–shell magnetic composites for highly efficient sorption of uranium (VI)

    Colloids Surf. A Physicochem. Eng. Asp.

    (2015)
  • C. Ding et al.

    Synergistic mechanism of U(VI) sequestration by magnetite-graphene oxide composites: evidence from spectroscopic and theoretical calculation

    Chem. Eng. J.

    (2017)
  • H. Gonzalez-Raymat et al.

    Unrefined humic substances as a potential low-cost amendment for the management of acidic groundwater contamination

    J. Environ. Manage.

    (2018)
  • H. Niu et al.

    Humic acid coated Fe3O4 magnetic nanoparticles as highly efficient Fenton-like catalyst for complete mineralization of sulfathiazole

    J. Hazard. Mater.

    (2011)
  • P. Singhal et al.

    Rapid extraction of uranium from sea water using Fe3O4 and humic acid coated Fe3O4 nanoparticles

    J. Hazard. Mater.

    (2017)
  • G. Dublet et al.

    Partitioning of uranyl between ferrihydrite and humic substances at acidic and circum-neutral pH

    Geochim. Cosmochim. Acta.

    (2017)
  • T. Missana et al.

    Experimental and modeling study of the uranium (VI) sorption on goethite

    J. Colloid Interface Sci.

    (2003)
  • H. Guo et al.

    Modeling and EXAFS investigation of U(VI) sequestration on Fe3O4/PCMs composites

    Chem. Eng. J.

    (2019)
  • Y. Sun et al.

    A robust prediction of U(VI) sorption on Fe3O4/activated carbon composites with surface complexation model

    Environ. Res.

    (2020)
  • T. Thielemann et al.

    Lignite and hard coal: energy suppliers for world needs until the year 2100 - an outlook

    Int. J. Coal Geol.

    (2007)
  • M. Havelcová et al.

    Sorption of metal ions on lignite and the derived humic substances

    J. Hazard. Mater.

    (2009)
  • J.B. Fein et al.

    Potentiometric titrations of Bacillus subtilis cells to low pH and a comparison of modeling approaches

    Geochim. Cosmochim. Acta

    (2005)
  • R. Zhang et al.

    Investigation of interaction between U(VI) and carbonaceous nanofibers by batch experiments and modeling study

    J. Colloid Interface Sci.

    (2015)
  • T. Hu et al.

    Application of three surface complexation models on U(VI) adsorption onto graphene oxide

    Chem. Eng. J.

    (2016)
  • K. Fukushi et al.

    A surface complexation model for sulfate and selenate on iron oxides consistent with spectroscopic and theoretical molecular evidence

    Geochim. Cosmochim. Acta.

    (2007)
  • Z. Chen et al.

    Synthesis of magnetic Fe3O4/CFA composites for the efficient removal of U(VI) from wastewater

    Chem. Eng. J.

    (2017)
  • E. Pehlivan et al.

    Comparison of adsorption capacity of young brown coals and humic acids prepared from different coal mines in Anatolia

    J. Hazard. Mater.

    (2006)
  • L. Peng et al.

    Modifying Fe3O4 nanoparticles with humic acid for removal of Rhodamine B in water

    J. Hazard. Mater.

    (2012)
  • C.J. Milne et al.

    Analysis of proton binding by a peat humic acid using a simple electrostatic model

    Geochim. Cosmochim. Acta

    (1995)
  • Cited by (19)

    • Uranium in natural waters and the environment: Distribution, speciation and impact

      2023, Applied Geochemistry
      Citation Excerpt :

      It includes both chemical speciation and reactive transport, and is fast. Experimentally, the interaction of metal ions with HS has often been studied in the presence of minerals, such as ferrihydrite or clay (Beneš et al., 1998; Dublet et al., 2017; Evans et al., 2011; Ho and Miller, 1985; Ivanov et al., 2012; Joseph et al., 2011; Lenhart and Honeyman, 1999; Payne et al., 1996; Schmeide et al., 2000; Zhang et al., 2021; Zhiwei et al., 2009). This addresses the question, does the presence of HS increase or decrease the binding of the metal ion to the solid phase?

    • Magnetic molecularly imprinted polymer combined with solid-phase extraction for detection of kojic acid in cosmetic products

      2022, Microchemical Journal
      Citation Excerpt :

      Moreover, in the deconvolution spectrum of Fe 2p in Fe3O4 (Fig. S5A), two peaks located at 710.8 eV and 724.6 eV were observed, which corresponded to Fe 2p3/2 and Fe 2p1/2, respectively [36]. In the deconvolution spectrum of O 1 s in Fe3O4 (Fig. S5B), three peaks located at 530.1 eV, 531.4 eV, and 531.6 eV were observed, which corresponded to Fe–O, O–CO, and –OH groups, respectively [37]. These results further indicated the successfull preparation of Fe3O4 and existence of carboxyl groups on the surface of Fe3O4.

    View all citing articles on Scopus
    View full text