Elsevier

Chemosphere

Volume 261, December 2020, 127704
Chemosphere

Rapid destruction of triclosan by Iron(III)-Tetraamidomacrocyclic ligand/hydrogen peroxide system

https://doi.org/10.1016/j.chemosphere.2020.127704Get rights and content

Highlights

  • Efficient removal of triclosan can be obtained by Fe(III)-TAML/H2O2 system.

  • The degradation of triclosan increases with the increasing pH.

  • NOM can cause slight inhibition for the degradation of triclosan.

  • Phenoxyl radical reaction and ring open reaction occurs.

  • Less antibiotic products are generated after triclosan degradation by Fe(III)-TAML/H2O2.

Abstract

Iron(III)-tetraamidomacrocyclic ligand (Fe(III)-TAML) activators can activate hydrogen peroxide to oxidize many kinds of organic pollutants. In this study, we investigated the degradation of triclosan, a widely used broad-spectrum bactericide, under the treatment of Fe(III)-TAML/H2O2 system at different pH conditions. We also studied the influence of natural organic matter (NOM) on the degradation process. Our results showed that complete removal of triclosan could be obtained within several minutes under the optimal conditions. The degradation of triclosan by Fe(III)-TAML/H2O2 system exhibited strong pH-dependence and the degradation rate increased with the increase in pH level from 7.0 to 10.0. When adding fulvic acid (FA) or humic acid (HA) in the reaction system, the degradation of triclosan could be suppressed slightly, and HA exhibited stronger inhibition than FA. Based on the analysis of reaction intermediates, phenoxyl radical reaction and ring open reaction were involved in the decomposition of triclosan. Significant inhibition of overall toxicity to Photobacterium phosphoreum further confirmed the high efficiency of Fe(III)-TAML/H2O2 system for the removal of antibiotic activities resulting from the parent triclosan molecule and its degradation products.

Introduction

Triclosan (5-chloro-2-(2,4-dichlorophenoxy)phenol) is primarily employed as the antibacterial, antimicrobial and preservative agent in consumer products and also as an active ingredient in personal care products, e.g., soaps, toothpaste, skin creams and deodorants (Sanchez-Prado et al., 2006; Aranami and Readman, 2007). However, due to the inability of conventional wastewater treatment processes (McAvoy et al., 2002; Sabaliunas et al., 2003; Winkler et al., 2007; Ying et al., 2009), triclosan cannot be removed efficiently, thus leading to the detection of its residue and derivatives in various environmental matrices, such as surface waters, sludge and sediments (Singer et al., 2002; Stackelberg et al., 2004; Weigel et al., 2004; Hua et al., 2005). As indicated by the results of a reconnaissance study by the United States Geological Survey, triclosan was confirmed to be one of the most frequently detected organic wastewater contaminants in 139 streams located in the USA, with the maximum concentration of 2.3 μg L−1 (Kolpin et al., 2002). Triclosan has exhibited acute and chronic toxicity to various aquatic organisms, including algae, photobacterium and fish, with the median effect concentration in the range of μg L−1 to mg L−1 (Adolfsson-Erici et al., 2002; Orvos et al., 2002). Moreover, Oliveira et al. (2009) found that triclosan could induce endocrine disruption for embryos and larvae of zebrafish, as indicated by the observations of hatching delay, teratogenic responses and mortality. Triclosan was also reported to bioaccumulate in kinds of aqueous species, e.g., algae, snails, fish and marine mammals (Lindstrom et al., 2002; Fernandes et al., 2011; Zhang et al., 2015), which might exhibit severe health effects along the food chain (Houtman et al., 2004; Coogan and La Point, 2008; Fair et al., 2009). In addition to the hazard of parent molecule, products of triclosan degradation were reported more toxic, which could be attributed to the generation of carcinogenic byproducts. Photodegradation of triclosan produces 2,8-dichlorodibenzo-p-dioxin, which is well recognized as a kind of carcinogen (Latch et al., 2005; Sanchez-Prado et al., 2006; Aranami and Readman, 2007; Wong-Wah-Chung et al., 2007). Biological methylation of triclosan generates more lipophilic and bioaccumulative methyl triclosan (Poiger et al., 2003). Moreover, triclosan reacting with free chlorine in wastewater and drinking water treatment would produce chloroform and chlorinated phenols, such as 2,4-dichlorophenol and 2,4,6-trichlorophenol (Rule et al., 2005), which are ranked by the US EPA as probable human carcinogens and listed on the Contaminant Candidate List (Song et al., 2012).

Considering the persistence and potential risk of triclosan in aquatic environment, it is necessary to develop efficient techniques to degrade it. It was reported that triclosan could not be effectively eliminated by traditional techniques, such as biofilm metabolism, chlorine disinfection, MnO2 oxidation and UV irradiation, and the toxic and carcinogenic byproducts, e.g., chloroform, chlorinated phenols and even dioxin-type intermediates were generated during the degradation process (Latch et al., 2003; Zhang and Huang, 2003; Rule et al., 2005; Ricart et al., 2010; Song et al., 2012). Recently, some advanced strategies, such as electrochemical (Martín de Vidales et al., 2013), sonochemical (Sanchez-Prado et al., 2008) and sonoelectrochemical (Ren et al., 2014) processes, have been developed to degrade triclosan, whereas high energy cost inevitably limits their factual application. In addition, Fenton-like techniques have been confirmed to efficiently deplete triclosan, while formation of ferric sludge and carcinogenic byproducts (e.g., 2,4-dichlorophenol) would still be problematic issues (Munoz et al., 2012; Song et al., 2012).

Iron(III)-tetraamidomacrocyclic ligand (Fe(III)-TAML) activators are regarded as functional analog of peroxidase enzyme (Sen Gupta et al., 2002). This kind of activators can catalyze hydrogen peroxide (H2O2) to oxidize substrates efficiently (Su et al., 2018). When Fe(III)-TAML reacts with H2O2, a high valence iron-oxo complex is formed and restores itself to Fe(III)-TAML while oxidizing substrates (Su et al., 2018). The reaction of Fe(III)-TAML/H2O2 system and substrates is well-known to be pH-dependent (Popescu et al., 2008; Wang et al., 2017a). Because the proportion of ionized Fe(III)-TAML species (i.e., [Fe(III)-TAML(OH)]2-) increases as pH increases. The [Fe(III)-TAML(OH)]2- form was reported to exhibit greater potential to be oxidized by H2O2 than [Fe(III)-TAML(OH2)]-, indicating that active Fe-TAML (e.g., Fe(IV)-TAML or Fe(V)-TAML) was more facile to be obtained at higher pH levels (Popescu et al., 2008; Wang et al., 2017a). Therefore, the reactivity of Fe(III)-TAML is enhanced dramatically with increasing pH. Fe(III)-TAML/H2O2 system can degrade varieties of contaminants such as phenols (Sen Gupta et al., 2002; Kundu et al., 2015; Wang et al., 2017a), dyes (Chahbane et al., 2007; Ghosh et al., 2008; Ellis et al., 2009), estrogens (Shappell et al., 2008; Chen et al., 2012; Onundi et al., 2017), organophosphorus pesticides (Chanda et al., 2006), molluscicides (Tang et al., 2016), pharmaceutical ingredients (Shen et al., 2011) and explosives (Kundu et al., 2013). The complete removal of pollutants can be obtained within few seconds at optimal conditions, in which trace concentration of Fe(III)-TAML is utilized (Wang et al., 2017a). Moreover, the previous study showed that the toxicity of contaminated water dramatically decreased after Fe(III)-TAML/H2O2 treatment, meanwhile no adverse effects on fish or microorganism were observed with the residual Fe(III)-TAML (Ellis et al., 2010). Therefore, Fe(III)-TAML/H2O2 exhibits great potential to accomplish high degradation efficiency for triclosan.

Natural organic matter (NOM) exists widely in water environment (Matilainen and Sillanpaa, 2010). When applying techniques to remove pollutants in water, we should take NOM into consideration, as it has tendency to interfere the process. Lado Ribeiro et al. (2019) have summarized the influence of organic matter in some different advanced oxidation technologies. For photolysis, photocatalysis, H2O2- and Fenton-based treatment, organic matter can cause both promotion and inhibition; while for O3- based water treatment methods, organic matter can only inhibit the procedure. However, for Fe(III)-TAML/H2O2 system, there has been no research about the influence of NOM. The objective of this study is to investigate the degradation efficiency of triclosan by Fe(III)-TAML/H2O2 system and the effects of pH and NOM on the reaction rates. Moreover, the underlying reaction mechanism and degradation pathway are elucidated based on the results of mass spectroscopy and theoretical calculation. Furthermore, the acute toxicity is also evaluated during degradation process.

Section snippets

Materials

Triclosan, sodium hydroxide (NaOH), perchloric acid (HClO4), sodium bicarbonate (NaHCO3), sodium carbonate (Na2CO3), disodium hydrogen phosphate (Na2HPO4), sodium dihydrogen phosphate (NaH2PO4), sodium sulfite (Na2SO3), sodium chloride (NaCl), HPLC grade methanol and acetonitrile purchased from Sigma-Aldrich. H2O2 (30%, w/w), prototype Fe(III)-TAML (Fig. 1) and NOM (i.e., humic acid (HA) and fulvic acid (FA)) were purchased from Fisher Scientific, GreenOx Catalysts (Mellon Institute,

Degradation of triclosan by Fe(III)-TAML/H2O2 system at different pHs

As shown in Fig. S1, when only adding Fe(III)-TAML or H2O2 to triclosan solution, the concentration of triclosan almost did not decrease in 10 min. That means if triclosan could decrease with both Fe(III)-TAML and H2O2, it is because of the cooperation between the two substances. As shown in Fig. 2, the degradation of triclosan by Fe(III)-TAML/H2O2 system exhibited strong pH-dependence, and the reaction rate increased with the increase in pH level. Almost complete degradation of triclosan by

Conclusions

In this study, we investigated the degradation of triclosan by Fe(III)-TAML/H2O2 system and the effects of pH and NOM. Our results showed that the degradation of triclosan by Fe(III)-TAML/H2O2 was highly pH dependent with the highest reaction rate at pH 10.0. In addition, the presence of NOM showed inhibition effects for the degradation of triclosan, and HA exhibited greater inhibition than FA at pH 8.0 and 9.0. However, the inhibition effect became insignificant when pH increased to 10.0,

Author statement

Sijia Liang: Methodology, Data curation, Writing-Reviewing and Editing, Zeyu Xian: Investigation, Haotian Yang: Investigation, Ziyu Wang: Investigation, Chao Wang: Conceptualization, Methodology, Data curation, Writing-Reviewing and Editing, Xiaoxia Shi: Investigation, Haoting Tian: Conceptualization, Writing-Reviewing

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 reported in this paper.

Acknowledgements

This work was financially supported by the National Key Research and Development Plans of Special Project for Site Soil (No. 2018YFC1802003), the National Science Foundation of China (No. 21906079), the Collaborative Innovation Center for Regional Environmental Quality, and International Institute for Environmental Studies. We thank the Analytical Center of Nanjing University for the characterization of samples and computational study.

References (72)

  • D.E. Latch et al.

    Photochemical conversion of triclosan to 2,8-dichlorodibenzo-p-dioxin in aqueous solution

    J. Photochem. Photobiol. Chem.

    (2003)
  • Y. Lei et al.

    Radical chemistry of diethyl phthalate oxidation via UV/peroxymonosulfate process: roles of primary and secondary radicals

    Chem. Eng. J.

    (2020)
  • H. Li et al.

    Peroxymonosulfate activation by iron(III)-tetraamidomacrocyclic ligand for degradation of organic pollutants via high-valent iron-oxo complex

    Water Res.

    (2018)
  • C. Luo et al.

    Oxidation of the odorous compound 2,4,6-trichloroanisole by UV activated persulfate: kinetics, products, and pathways

    Water Res.

    (2016)
  • J. Ma et al.

    Impacts of inorganic anions and natural organic matter on thermally activated persulfate oxidation of BTEX in water

    Chemosphere

    (2018)
  • A. Matilainen et al.

    Removal of natural organic matter from drinking water by advanced oxidation processes

    Chemosphere

    (2010)
  • M. Munoz et al.

    Triclosan breakdown by Fenton-like oxidation

    Chem. Eng. J.

    (2012)
  • B. Quan et al.

    Technology and principle of removing triclosan from aqueous media: a review

    Chem. Eng. J.

    (2019)
  • Y.-Z. Ren et al.

    Sonoelectrochemical degradation of triclosan in water

    Ultrason. Sonochem.

    (2014)
  • M. Ricart et al.

    Triclosan persistence through wastewater treatment plants and its potential toxic effects on river biofilms

    Aquat. Toxicol.

    (2010)
  • D. Sabaliunas et al.

    Environmental fate of triclosan in the river aire basin, UK

    Water Res.

    (2003)
  • L. Sanchez-Prado et al.

    Sonochemical degradation of triclosan in water and wastewater

    Ultrason. Sonochem.

    (2008)
  • L. Sanchez-Prado et al.

    Monitoring the photochemical degradation of triclosan in wastewater by UV light and sunlight using solid-phase microextraction

    Chemosphere

    (2006)
  • N.S. Shah et al.

    Hydroxyl and sulfate radical mediated degradation of ciprofloxacin using nano zerovalent manganese catalyzed S2O82−

    Chem. Eng. J.

    (2019)
  • Z. Song et al.

    Efficient oxidative degradation of triclosan by using an enhanced Fenton-like process

    Chem. Eng. J.

    (2012)
  • P.E. Stackelberg et al.

    Persistence of pharmaceutical compounds and other organic wastewater contaminants in a conventional drinking-water-treatment plant

    Sci. Total Environ.

    (2004)
  • H. Su et al.

    Quantitative structure-activity relationship for the oxidation of aromatic organic contaminants in water by TAML/H2O2

    Water Res.

    (2018)
  • O. Tantawi et al.

    A rapid and economical method for the quantification of hydrogen peroxide (H2O2) using a modified HPLC apparatus

    Sci. Total Environ.

    (2019)
  • C. Wang et al.

    Rapid destruction of Tetrabromobisphenol A by Iron(III)-Tetraamidomacrocyclic ligand/layered double hydroxide composite/H2O2 System

    Environ. Sci. Technol.

    (2017)
  • S. Wang et al.

    Enhanced biodegradation of triclosan by means of gamma irradiation

    Chemosphere

    (2017)
  • S. Weigel et al.

    Determination of selected pharmaceuticals and caffeine in sewage and seawater from Tromso/Norway with emphasis on ibuprofen and its metabolites

    Chemosphere

    (2004)
  • P. Wong-Wah-Chung et al.

    Photochemical behaviour of triclosan in aqueous solutions: kinetic and analytical studies

    J. Photochem. Photobiol. Chem.

    (2007)
  • Q. Wu et al.

    Oxidative removal of selected endocrine-disruptors and pharmaceuticals in drinking water treatment systems, and identification of degradation products of triclosan

    Sci. Total Environ.

    (2012)
  • B. Yang et al.

    Oxidation of triclosan by ferrate: reaction kinetics, products identification and toxicity evaluation

    J. Hazard Mater.

    (2011)
  • J.C. Yu et al.

    Photocatalytic oxidation of triclosan

    Chemosphere

    (2006)
  • Z. Zhang et al.

    Enhanced dechlorination of triclosan by hydrated electron reduction in aqueous solution

    Chem. Eng. J.

    (2015)
  • Cited by (7)

    • Preparations of cyclodextrin polymer and MgO jointly entrapped iron(III)-TAML catalysts for the removal of aromatic pollutants in water

      2023, Separation and Purification Technology
      Citation Excerpt :

      Consequently, naturally occurring DOM at environmentally relevant levels are found to pose no interference on Fe(III)-TAML/H2O2 [16,24,25]. Nonetheless, the reactivity of Fe(III)-TAML/H2O2 is inhibited in water containing high levels of DOM (dozens of mg/L) and/or artificially isolated DOM (e.g., humic acids) [26–28]. Growing studies have highlighted the concerns about the recycling, reactivity and applicability of Fe(III)-TAML in water [20,29].

    • Emerging Organic Contaminants in Chinese Surface Water: Identification of Priority Pollutants

      2022, Engineering
      Citation Excerpt :

      Terbutryn, a triazine herbicide used to control broadleaf weeds, free-floating weeds, and algae, was banned in EU agriculture in 2002 and recognized as a priority substance (Directive 2013/39/EU) due to its bioaccumulation potential [103,104]. Triclosan is a widely used antibacterial preservative, which undergoes biomagnification, exhibits potential endocrine disrupting effects, and poses acute and chronic toxicity to various organisms [105]. For these reasons, the use of triclosan in soap and other personal care products has been banned by the US FDA [106]; furthermore, triclosan has been identified as a priority substance (NORMAN Category 1)† [19].

    • Photodegradation of Bisphenol A and Triclosan using ZnCl<inf>2</inf>(6OMeQ)<inf>2</inf> complex as photosensitizer under homogeneous conditions

      2021, Results in Engineering
      Citation Excerpt :

      Among the multiple applications of the organometallic photosensitizer as catalysts in chemical reactions [12–14] their capacity has also been demonstrated for photosensitizer applications in water treatment, whereby the photosensitive material is fixed on a polymer surface in solid phase, studies reveal that using polymers such as silicone preserve the fluorescence signals so it is very convenient to fix it on these materials [32]. Several researchers have reported significant results in terms of degradation efficiency for the treatment of BPA [[[33 - 36]]] and TCS [[[37 - 40]]], however, in this research, aims to demonstrate the participation of the singlet molecular oxygen and its application against new photosensitizing compounds. Additionally, it has been established that the complexes ZnCl2(6OMeQ)2 have a good quantum yield for the generation of the singlet molecular oxygen [23], therefore, this research generates important contributions for its application in environmental matrices.

    • Green synthesis of lignin nanorods/g-C<inf>3</inf>N<inf>4</inf> nanocomposite materials for efficient photocatalytic degradation of triclosan in environmental water

      2021, Chemosphere
      Citation Excerpt :

      Fig. 4d displayed the photodegradation rate of TCS with respect to various photodegradation times. It may be noticed from Fig. 4d that photo-catalytic degradation rate of TCS enhanced with a raise in pH range due to less surface adsorption of TCS onto the surface of nanocomposite photocatalyst that promotes photodegradation at higher pH (Liang et al., 2020; Jiang et al., 2017). Moreover, excess OH− ions in alkaline conditions may also react with the photo-generated holes (h+) to favor the creation of more OH, which significantly improves the photocatalytic degradation of the TCS (Liang et al., 2020: Wang et al., 2008).

    View all citing articles on Scopus
    View full text