[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (59)

Search Parameters:
Keywords = iron persulfate

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
32 pages, 7907 KiB  
Article
Degradation Efficiency and Mechanism of Tetracycline in Water by Activated Persulfate Using Biochar-Loaded Nano Zero-Valent Iron
by Bojiao Yan, Xueqi Li, Xiaoyan Wang, Ping Yang, Hai Lu and Xiaoyu Zhang
Molecules 2024, 29(16), 3875; https://doi.org/10.3390/molecules29163875 - 15 Aug 2024
Viewed by 745
Abstract
Tetracycline (TC) contamination in water is one of the key issues in global environmental protection, and traditional water treatment methods are difficult to remove antibiotic pollutants.Therefore, efficient and environmentally friendly treatment technologies are urgently needed. In this study, activated persulfate (PS) using a [...] Read more.
Tetracycline (TC) contamination in water is one of the key issues in global environmental protection, and traditional water treatment methods are difficult to remove antibiotic pollutants.Therefore, efficient and environmentally friendly treatment technologies are urgently needed. In this study, activated persulfate (PS) using a biochar-loaded nano zero-valent iron (BC-nZVI) advanced oxidation system was used to investigate the degradation effect, influencing factors, and mechanism of TC. BC-nZVI was prepared using the liquid-phase reduction method, and its structure and properties were analyzed by various characterization means. The results showed that nZVI was uniformly distributed on the surface or in the pores of BC, forming a stable complex. Degradation experiments showed that the BC-nZVI/PS system could degrade TC up to 99.57% under optimal conditions. The experiments under different conditions revealed that the iron-carbon ratio, dosing amount, PS concentration, and pH value all affected the degradation efficiency. Free radical burst and electron paramagnetic resonance (EPR) experiments confirmed the dominant roles of hydroxyl and sulfate radicals in the degradation process, and LC–MS experiments revealed the multi-step reaction process of TC degradation. This study provides a scientific basis for the efficient treatment of TC pollution in water. Full article
Show Figures

Figure 1

Figure 1
<p>TC standard curve.</p>
Full article ">Figure 2
<p>SEM-EDS diagram of BC (<b>a</b>), nZVI (<b>b</b>), BC-nZVI (new) (<b>c</b>), and BC-nZVI (used) (<b>d</b>).</p>
Full article ">Figure 2 Cont.
<p>SEM-EDS diagram of BC (<b>a</b>), nZVI (<b>b</b>), BC-nZVI (new) (<b>c</b>), and BC-nZVI (used) (<b>d</b>).</p>
Full article ">Figure 3
<p>XRD results of BC, nZVI, BC-nZVI (new), and BC-nZVI (used).</p>
Full article ">Figure 4
<p>N<sub>2</sub> adsorption–desorption isotherms of different materials.</p>
Full article ">Figure 5
<p>XPS diagram of different materials.</p>
Full article ">Figure 6
<p>FT-IR diagram of different materials.</p>
Full article ">Figure 7
<p>The degradation results of TC under different systems.</p>
Full article ">Figure 8
<p>The degradation results of TC under different iron-carbon ratios.</p>
Full article ">Figure 9
<p>The degradation results of TC under different BC-nZVI dosages.</p>
Full article ">Figure 10
<p>The degradation results of TC at different initial concentrations.</p>
Full article ">Figure 11
<p>The degradation results of TC at different PS concentrations.</p>
Full article ">Figure 12
<p>The degradation results of TC at different pH values.</p>
Full article ">Figure 13
<p>The influence of HCO<sub>3</sub><sup>−</sup> on the degradation of TC.</p>
Full article ">Figure 14
<p>The influence of Cl<sup>−</sup> on the degradation of TC.</p>
Full article ">Figure 15
<p>The influence of SO<sub>4</sub><sup>2−</sup> on the degradation of TC.</p>
Full article ">Figure 16
<p>Degradation effect of BC-nZVI/PS system on TC in raw water.</p>
Full article ">Figure 17
<p>EPR detection spectrum.</p>
Full article ">Figure 18
<p>XPS detection spectrum.</p>
Full article ">Figure 19
<p>Reaction mechanisms of BC-nZVI/PS system [<a href="#B55-molecules-29-03875" class="html-bibr">55</a>].</p>
Full article ">Figure 20
<p>Spectral diagram of TC degradation products. (<b>a</b>), The reaction time is 0 min (<b>b</b>), The reaction time is 120 min (<b>c</b>), The reaction time is 5 min (<b>d</b>), The reaction time is 15 min (<b>e</b>), The reaction time is 30 min (<b>f</b>), The reaction time is 60 min.</p>
Full article ">Figure 20 Cont.
<p>Spectral diagram of TC degradation products. (<b>a</b>), The reaction time is 0 min (<b>b</b>), The reaction time is 120 min (<b>c</b>), The reaction time is 5 min (<b>d</b>), The reaction time is 15 min (<b>e</b>), The reaction time is 30 min (<b>f</b>), The reaction time is 60 min.</p>
Full article ">Figure 20 Cont.
<p>Spectral diagram of TC degradation products. (<b>a</b>), The reaction time is 0 min (<b>b</b>), The reaction time is 120 min (<b>c</b>), The reaction time is 5 min (<b>d</b>), The reaction time is 15 min (<b>e</b>), The reaction time is 30 min (<b>f</b>), The reaction time is 60 min.</p>
Full article ">Figure 21
<p>Speculated pathways for TC oxidation degradation.</p>
Full article ">Figure 22
<p>Physical diagram of main experimental equipment. (<b>a</b>), Vacuum drying oven (<b>b</b>), Water-bath constant tem-perature shaker.</p>
Full article ">Figure 23
<p>Schematic diagram of preparation of biochar-supported nano zero-valent iron by liquid-phase reduction method.</p>
Full article ">
13 pages, 1865 KiB  
Article
A Novel Method for the Highly Effective Removal of Binary Dyes from Colored Dyeing Wastewater by Periodic Reversal/Direct Current-Activated Persulfate
by Zhaonan Sun, Wenjie Ren, Ke Shi, Wei Kou and Yujie Feng
Sustainability 2024, 16(10), 4057; https://doi.org/10.3390/su16104057 - 13 May 2024
Viewed by 709
Abstract
In recent years, electrochemical synergistic activation of persulfate (PDS) degradation technology has demonstrated significant potential in wastewater treatment applications. Given the challenges posed by the complex water quality, high COD content, and recalcitrant degradation of dyeing wastewater, this study aimed to evaluate the [...] Read more.
In recent years, electrochemical synergistic activation of persulfate (PDS) degradation technology has demonstrated significant potential in wastewater treatment applications. Given the challenges posed by the complex water quality, high COD content, and recalcitrant degradation of dyeing wastewater, this study aimed to evaluate the efficacy of iron/aluminum dual-electrode electrochemical activation of PDS for degrading simulated dyeing wastewater. The results showed that under optimal conditions, utilizing both periodic reversal and direct current electrochemical activation of PDS achieved removal rates of 99.2% and 98.3% for Reactive Black 5 (RB5) and Reactive Red X-3B (RRX-3B), respectively, demonstrating promising removal efficiency. Notably, the removal efficiency of RB5 surpassed that of RRX-3B, suggesting a dependence on initial concentration influencing reaction kinetics. Furthermore, full-spectrum scanning and quenching experiments revealed that RB5 and RRX-3B were primarily degraded through the potent oxidation action of SO4· and ·OH, with a small number of intermediates present in the solution. Periodic reversal proved effective in mitigating electrode passivation and enhancing electrode longevity. This study provides a highly effective removal method of binary dyes from dyeing wastewater by periodic reversal Fe-Al dual-electrode electrochemical activation of PDS technology, offering valuable insights for sustainable treatment of dyeing wastewater with binary components. Full article
(This article belongs to the Section Pollution Prevention, Mitigation and Sustainability)
Show Figures

Figure 1

Figure 1
<p>UV-VIS spectral scanning of dyeing wastewater with binary components, RB5 and RRX-3B.</p>
Full article ">Figure 2
<p>Influence of reaction voltage on the removal effect of binary dyes in wastewater (<b>a</b>) and DC (<b>b</b>) periodically reversing.</p>
Full article ">Figure 3
<p>Influence of PDS dosage on the removal effect of dyes in wastewater with binary components: (<b>a</b>) DC; (<b>b</b>) periodically reversing.</p>
Full article ">Figure 4
<p>Influence of pad spacing on the removal effect of dyes in wastewater with binary components: (<b>a</b>) DC; (<b>b</b>) periodically reversing.</p>
Full article ">Figure 5
<p>Influence of initial pH value on the removal effect of dyes in wastewater with binary components: (<b>a</b>) DC; (<b>b</b>) periodically reversing.</p>
Full article ">Figure 6
<p>Influence of periodically reversing/DC-assisted process on residual PDS.</p>
Full article ">Figure 7
<p>Influence of quenching agent on the removal efficiency ((<b>a</b>) direct current; (<b>b</b>) periodically reversing).</p>
Full article ">Figure 8
<p>UV-VIS spectral scanning ((<b>a</b>) Periodically reversing, (<b>b</b>) DC).</p>
Full article ">
22 pages, 9063 KiB  
Article
Remediation of Polycyclic Aromatic Hydrocarbon-Contaminated Soil by Using Activated Persulfate with Carbonylated Activated Carbon Supported Nanoscale Zero-Valent Iron
by Changzhao Chen, Zhe Yuan, Shenshen Sun, Jiacai Xie, Kunfeng Zhang, Yuanzheng Zhai, Rui Zuo, Erping Bi, Yufang Tao and Quanwei Song
Catalysts 2024, 14(5), 311; https://doi.org/10.3390/catal14050311 - 8 May 2024
Cited by 2 | Viewed by 1103
Abstract
Soil contamination by polycyclic aromatic hydrocarbons (PAHs) has been an environmental issue worldwide, which aggravates the ecological risks faced by animals, plants, and humans. In this work, the composites of nanoscale zero-valent iron supported on carbonylated activated carbon (nZVI-CAC) were prepared and applied [...] Read more.
Soil contamination by polycyclic aromatic hydrocarbons (PAHs) has been an environmental issue worldwide, which aggravates the ecological risks faced by animals, plants, and humans. In this work, the composites of nanoscale zero-valent iron supported on carbonylated activated carbon (nZVI-CAC) were prepared and applied to activate persulfate (PS) for the degradation of PAHs in contaminated soil. The prepared nZVI-CAC catalyst was characterized by scanning electron microscopy (SEM), X-ray diffractometer (XRD), Fourier transform infrared spectroscopy (FTIR), and X-ray photoelectron spectroscopy (XPS). It was found that the PS/nZVI-CAC system was superior for phenanthrene (PHE) oxidation than other processes using different oxidants (PS/nZVI-CAC > PMS/nZVI-CAC > H2O2/nZVI-CAC) and it was also efficient for the degradation of other six PAHs with different structures and molar weights. Under optimal conditions, the lowest and highest degradation efficiencies for the selected PAHs were 60.8% and 90.7%, respectively. Active SO4−• and HO were found to be generated on the surface of the catalysts, and SO4−• was dominant for PHE oxidation through quenching experiments. The results demonstrated that the heterogeneous process using activated PS with nZVI-CAC was effective for PAH degradation, which could provide a theoretical basis for the remediation of PAH-polluted soil. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD pattern of CAC and nZVI-CAC; (<b>b</b>) FTIR spectra of CAC and nZVI-CAC before and after the reaction.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM of CAC; (<b>b</b>,<b>c</b>) SEM of nZVI-CAC; (<b>d</b>–<b>h</b>) TEM of nZVI-CAC; (<b>i</b>–<b>l</b>) EDS of nZVI-CAC.</p>
Full article ">Figure 3
<p>The adsorption–desorption isotherms and the pore size distribution (the insets of figure) of CAC (<b>a</b>) and nZVI-CAC (<b>b</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) The full-scale XPS spectrum; (<b>b</b>) the C 1s XPS analysis; (<b>c</b>) the O 1s XPS analysis of CAC and nZVI-CAC; (<b>d</b>) the Fe 2p analysis of nZVI-CAC.</p>
Full article ">Figure 5
<p>The greenness of the analysis method.</p>
Full article ">Figure 6
<p>(<b>a</b>) PHE oxidation in different systems; (<b>b</b>) pseudo first order kinetic fit under optimal conditions; (<b>c</b>) PHE oxidation by using different oxidants. Conditions: except for the tested factor, the others were fixed, oxidant = 100 mmol kg<sup>−1</sup>, [Catalyst] = 10 g kg<sup>−1</sup>.</p>
Full article ">Figure 7
<p>(<b>a</b>) The effects of mass ratio of Fe/CAC; (<b>b</b>) the nZVI-CAC dosage on PHE oxidation; (<b>c</b>) PS concentration on PHE oxidation. Except for the tested factor, the others were fixed, Fe/CAC = 1:1, [PS] = 100 mmol kg<sup>−1</sup>, [Catalyst] = 10 g kg<sup>−1</sup>.</p>
Full article ">Figure 8
<p>Oxidation of various PAHs by the system of PS and PS/nZVI-CAC. Conditions: [PS] = 200 mmol kg<sup>−1</sup>, [Catalyst] = 10 g kg<sup>−1</sup>.</p>
Full article ">Figure 9
<p>The effects of radical scavengers on PHE oxidation; conditions: [PS] = 200 mmol/kg, [nZVI-CAC] = 10 g kg<sup>−1</sup>, [Scavenger] = 500 mmol kg<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Mechanism of PS/nZVI-CAC system for PHE oxidation.</p>
Full article ">
16 pages, 2050 KiB  
Article
Novel Oxidation Strategies for the In Situ Remediation of Chlorinated Solvents from Groundwater—A Bench-Scale Study
by Alicia Cano-López, Lidia Fernandez-Rojo, Leónidas Pérez-Estrada, Sònia Jou-Claus, Marta Batriu, Carme Bosch, Xavier Martínez-Lladó, Joana Baeta Trias, Ricard Mora Vilamaña, Mònica Escolà Casas and Víctor Matamoros
Water 2024, 16(9), 1241; https://doi.org/10.3390/w16091241 - 26 Apr 2024
Viewed by 1230
Abstract
Industrial chlorinated solvents continue to be among the most significant issues in groundwater (GW) pollution worldwide. This study assesses the effectiveness of eight novel oxidation treatments, including persulfate (PS), ferrous sulfate, sulfidated nano-zero valent iron (S-nZVI), and potassium ferrate, along with their combinations, [...] Read more.
Industrial chlorinated solvents continue to be among the most significant issues in groundwater (GW) pollution worldwide. This study assesses the effectiveness of eight novel oxidation treatments, including persulfate (PS), ferrous sulfate, sulfidated nano-zero valent iron (S-nZVI), and potassium ferrate, along with their combinations, for the potential in situ remediation of GW polluted with chlorinated solvents (1,2-dichloroethylene, trichloroethylene, and tetrachloroethylene). Our bench-scale results reveal that the combined addition of PS and S-nZVI can effectively eliminate trichloroethylene (10 µg/L), achieving removal rates of up to 80% and 92% within 1 h, respectively, when using synthetic GW. In the case of real GW, this combination achieved removal rates of 69, 99, and 92% for cis-1,2-dichloroethylene, trichloroethylene, and tetrachloroethylene, respectively, within 24 h. Therefore, this proposed remediation solution resulted in a significant reduction in the environmental risk quotient, shifting it from a high-risk (1.1) to a low-risk (0.2) scenario. Furthermore, the absence of transformation products, such as vinyl chloride, suggests the suitability of employing this solution for the in situ remediation of GW polluted with chlorinated solvents. Full article
(This article belongs to the Special Issue New Technologies for Soil and Groundwater Remediation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Layout showing the groundwater (GW) sampling points (three monitoring wells (MW) and one groundwater well) and water flow direction.</p>
Full article ">Figure 2
<p>Comparison of eight novel treatment methods for the removal of TCE spiked to 100 µg/L in synthetic GW (Batch 1). Various concentrations of PS, Fe(II), and Fe(VI) (0, 1, and 10 mM) and S-nZVI (0, 1, and 10 g/L) were tested (Control, Low, and High, respectively) at different incubation times (1, 2.5, and 5 h). Significant differences between samples, identified by different lower-case letters, were determined by a Friedman test with FDR correction for concentrations and treatments (<span class="html-italic">p</span> &lt; 0.05) due to paired and non-parametric samples.</p>
Full article ">Figure 3
<p>Removal efficiencies of chlorinated solvents (DCE, TCE, and PCE) in GW exposed for different incubation times (0, 1, 2.5, and 5 h) and the best preselected oxidation treatments, at 10 mM, with each compound spiked at 10 µg/L (Batch 2). Significant differences between treatments within each compound, identified by different lower-case letters, were determined by a Friedman test with FDR correction for treatments (<span class="html-italic">p</span> &lt; 0.05) due to paired and non-parametric samples.</p>
Full article ">Figure 4
<p>Removal of chlorinated solvents (DCE, TCE, and PCE) by a PS + S-nZVI treatment after a 5 h incubation time in spiked (10 µg/L each compound) and unspiked real Besòs GW and 10 mM of oxidant (Batch 3). Significant differences between the spiked and unspiked reactions’ removals within each compound, identified by different lower-case letters, were determined by a Friedman test with FDR correction for treatments (<span class="html-italic">p</span> &lt; 0.05) due to paired and non-parametric samples. Since DCE was not detected in unspiked samples (0 nor 5 h), the statistical comparison had enough samples to be compared.</p>
Full article ">Figure 5
<p>Removal of chlorinated solvents by PS + S-nZVI treatment at 0, 9, and 24 h of incubation in real Besòs GW spiked with 10 µg/L of each of the selected compounds (DCE, TCE, and PCE) and 10 mM of oxidant (Batch 4). Significant differences between different incubation times within each compound, identified by different lower-case letters, were determined by a Friedman test with FDR correction for the treatments of each compound (<span class="html-italic">p</span> &lt; 0.05) due to paired and non-parametric samples.</p>
Full article ">
36 pages, 24439 KiB  
Review
Progress in the Elimination of Organic Contaminants in Wastewater by Activation Persulfate over Iron-Based Metal–Organic Frameworks
by Keke Zhi, Jiajun Xu, Shi Li, Lingjie Luo, Dong Liu, Zhe Li, Lianghui Guo and Junwei Hou
Nanomaterials 2024, 14(5), 473; https://doi.org/10.3390/nano14050473 - 5 Mar 2024
Viewed by 1545
Abstract
The release of organic contaminants has grown to be a major environmental concern and a threat to the ecology of water bodies. Persulfate-based Advanced Oxidation Technology (PAOT) is effective at eliminating hazardous pollutants and has an extensive spectrum of applications. Iron-based metal–organic frameworks [...] Read more.
The release of organic contaminants has grown to be a major environmental concern and a threat to the ecology of water bodies. Persulfate-based Advanced Oxidation Technology (PAOT) is effective at eliminating hazardous pollutants and has an extensive spectrum of applications. Iron-based metal–organic frameworks (Fe-MOFs) and their derivatives have exhibited great advantages in activating persulfate for wastewater treatment. In this article, we provide a comprehensive review of recent research progress on the significant potential of Fe-MOFs for removing antibiotics, organic dyes, phenols, and other contaminants from aqueous environments. Firstly, multiple approaches for preparing Fe-MOFs, including the MIL and ZIF series were introduced. Subsequently, removal performance of pollutants such as antibiotics of sulfonamides and tetracyclines (TC), organic dyes of rhodamine B (RhB) and acid orange 7 (AO7), phenols of phenol and bisphenol A (BPA) by various Fe-MOFs was compared. Finally, different degradation mechanisms, encompassing free radical degradation pathways and non-free radical degradation pathways were elucidated. This review explores the synthesis methods of Fe-MOFs and their application in removing organic pollutants from water bodies, providing insights for further refining the preparation of Fe-MOFs. Full article
(This article belongs to the Section Inorganic Materials and Metal-Organic Frameworks)
Show Figures

Figure 1

Figure 1
<p>The overview of strategies to synthesize MOF catalysts [<a href="#B36-nanomaterials-14-00473" class="html-bibr">36</a>]. (<b>a</b>) ‘ship-in-bottle’ and (<b>c</b>) ‘bottle-around-ship’ approaches both yielding (<b>b</b>). Adapted with permission from Ref. [<a href="#B36-nanomaterials-14-00473" class="html-bibr">36</a>]. Copyright 2022, copyright Kathrin L. Kollmannsberger.</p>
Full article ">Figure 2
<p>The SEM of MIL-53(Fe) [<a href="#B43-nanomaterials-14-00473" class="html-bibr">43</a>] (<b>a</b>), MIL-88(Fe) [<a href="#B44-nanomaterials-14-00473" class="html-bibr">44</a>] (<b>b</b>), MIL-101(Fe) [<a href="#B45-nanomaterials-14-00473" class="html-bibr">45</a>] (<b>c</b>), MIL-100 (Fe) [<a href="#B46-nanomaterials-14-00473" class="html-bibr">46</a>] (<b>d</b>). Adapted with permission from Ref. [<a href="#B43-nanomaterials-14-00473" class="html-bibr">43</a>]. Copyright 2017, copy-right Xuechuan Gao. Adapted with permission from Ref. [<a href="#B44-nanomaterials-14-00473" class="html-bibr">44</a>]. Copyright 2016, copy-right Xuechao Cai. Adapted with permission from Ref. [<a href="#B45-nanomaterials-14-00473" class="html-bibr">45</a>]. Copyright 2024, copy-right Abhijit Das.</p>
Full article ">Figure 3
<p>(<b>a</b>) The schematic diagram for MIL-100(Fe) synthesis hydrothermal method. (<b>b</b>) The SEM of MIL-100(Fe) [<a href="#B42-nanomaterials-14-00473" class="html-bibr">42</a>]. Adapted with permission from Ref. [<a href="#B42-nanomaterials-14-00473" class="html-bibr">42</a>]. Copyright 2024, copy-right Huizhong Zhao.</p>
Full article ">Figure 4
<p>(<b>a</b>) The schematic diagram for MIL synthesis procedures via microwave-assisted technique. (<b>b</b>,<b>c</b>) The SEM of MIL-88(Fe) [<a href="#B67-nanomaterials-14-00473" class="html-bibr">67</a>]. Adapted with permission from Ref. [<a href="#B67-nanomaterials-14-00473" class="html-bibr">67</a>]. Copyright 2022, copy-right Mahmoud Y. Zorainy.</p>
Full article ">Figure 5
<p>(<b>a</b>) The schematic diagram of catalysis synthesis by DGC. (<b>b</b>) The SEM of MIL-100(Fe)-DGC. (<b>c</b>) The SEM of MIL-100(Fe) [<a href="#B41-nanomaterials-14-00473" class="html-bibr">41</a>]. Adapted with permission from Ref. [<a href="#B41-nanomaterials-14-00473" class="html-bibr">41</a>]. Copyright 2019, copy-right Yanshu Luo.</p>
Full article ">Figure 6
<p>(<b>a</b>) The schematic diagram of catalysis synthesis by hydrothermal method. (<b>b</b>). The SEM of Fe-ZIF-8 [<a href="#B72-nanomaterials-14-00473" class="html-bibr">72</a>]. Adapted with permission from Ref. [<a href="#B72-nanomaterials-14-00473" class="html-bibr">72</a>]. Copyright 2023, copy-right Entisar M. Khudhair.</p>
Full article ">Figure 7
<p>The schematic diagram of the preparation of MIL-100(Fe)/Ti<sub>3</sub>C<sub>2</sub> [<a href="#B75-nanomaterials-14-00473" class="html-bibr">75</a>]. Adapted with permission from Ref. [<a href="#B75-nanomaterials-14-00473" class="html-bibr">75</a>]. Copyright 2019, copy-right Hanmei Wang.</p>
Full article ">Figure 8
<p>(<b>a</b>) The fabrication strategy of Fe<sub>3</sub>O<sub>4</sub>@MIL-100(Fe)/Ag nanocomposites. (<b>b</b>) The SEM of Fe<sub>3</sub>O<sub>4</sub>. (<b>c</b>) The SEM of Fe<sub>3</sub>O<sub>4</sub> [<a href="#B76-nanomaterials-14-00473" class="html-bibr">76</a>]. Adapted with permission from Ref. [<a href="#B76-nanomaterials-14-00473" class="html-bibr">76</a>]. Copyright 2020, copy-right Shuquan Chang.</p>
Full article ">Figure 9
<p>The fabrication strategy of Fe<sub>3</sub>O<sub>4</sub>@GO@MIL-101(Fe) [<a href="#B78-nanomaterials-14-00473" class="html-bibr">78</a>]. Adapted with permission from Ref. [<a href="#B78-nanomaterials-14-00473" class="html-bibr">78</a>]. Copyright 2024, copy-right Lin He.</p>
Full article ">Figure 10
<p>The schematic diagram of the preparation of CoFe-N-CNTs [<a href="#B79-nanomaterials-14-00473" class="html-bibr">79</a>]. Adapted with permission from Ref. [<a href="#B79-nanomaterials-14-00473" class="html-bibr">79</a>]. Copyright 2020, copy-right Zhen Wang.</p>
Full article ">Figure 11
<p>(<b>a</b>) The schematic preparation of kapok biomass-C@MOFs-C [<a href="#B82-nanomaterials-14-00473" class="html-bibr">82</a>]. (<b>b</b>) The schematic diagram showing the synthetic route for MPN@NH<sub>2</sub>-MIL-101(Fe) (ⅰ) is the steps of the experiment and (ⅱ) is the figures of reactants and products [<a href="#B83-nanomaterials-14-00473" class="html-bibr">83</a>]. (<b>c</b>) The schematic preparation of palm tree biochar MOF [<a href="#B81-nanomaterials-14-00473" class="html-bibr">81</a>]. Adapted with permission from Ref. [<a href="#B82-nanomaterials-14-00473" class="html-bibr">82</a>]. Copyright 2020, copy-right Yang Zhao. Adapted with permission from Ref. [<a href="#B83-nanomaterials-14-00473" class="html-bibr">83</a>]. Copyright 2023, copy-right Aaron Albert Aryee. Adapted with permission from Ref. [<a href="#B81-nanomaterials-14-00473" class="html-bibr">81</a>]. Copyright 2023, copy-right Hanane Chakhtouna.</p>
Full article ">Figure 12
<p>The process of molecular imprinting [<a href="#B92-nanomaterials-14-00473" class="html-bibr">92</a>]. Adapted with permission from Ref. [<a href="#B92-nanomaterials-14-00473" class="html-bibr">92</a>]. Copyright 2014, copy-right Jennifer E. Lofgreen.</p>
Full article ">Figure 13
<p>(<b>a</b>). The schematic preparation of MIPMIL100(Fe) [<a href="#B96-nanomaterials-14-00473" class="html-bibr">96</a>]. (<b>b</b>) The degradation of DEP from materials after adsorption [<a href="#B96-nanomaterials-14-00473" class="html-bibr">96</a>]. Adapted with permission from Ref. [<a href="#B96-nanomaterials-14-00473" class="html-bibr">96</a>]. Copyright 2020, copy-right Xitong Li.</p>
Full article ">Figure 14
<p>[<a href="#B106-nanomaterials-14-00473" class="html-bibr">106</a>] (<b>a</b>) The catalytic oxidation of SMX under different catalytic systems. [<a href="#B107-nanomaterials-14-00473" class="html-bibr">107</a>] (<b>b</b>) The selective adsorption and catalysis performance of NH<sub>2</sub>-MIL-53(Fe) and a series of MIP materials. [<a href="#B108-nanomaterials-14-00473" class="html-bibr">108</a>] (<b>c</b>) Various catalysts in PMS activation for SMX degradation. Adapted with permission from Ref. [<a href="#B106-nanomaterials-14-00473" class="html-bibr">106</a>]. Copyright 2023, copy-right Abdul Hannan Asif. Adapted with permission from Ref. [<a href="#B107-nanomaterials-14-00473" class="html-bibr">107</a>]. Copyright 2022, copy-right Yongchang Xie. Adapted with permission from Ref. [<a href="#B108-nanomaterials-14-00473" class="html-bibr">108</a>]. Copyright 2023, copy-right Liqin Chen.</p>
Full article ">Figure 15
<p>(<b>a</b>) Catalyst dosage effects of the degradation system on the TC degradation efficiency [<a href="#B118-nanomaterials-14-00473" class="html-bibr">118</a>]. (<b>b</b>) The degradation of TC by PS only, Fe-MOF/PS, Fe/N-MOF/PS, Co/N-MOF/PS and FeCo/N-MOF (7:3)/PS [<a href="#B119-nanomaterials-14-00473" class="html-bibr">119</a>]. (<b>c</b>). The comparison of different systems for TC removal [<a href="#B83-nanomaterials-14-00473" class="html-bibr">83</a>]. Adapted with permission from Ref. [<a href="#B118-nanomaterials-14-00473" class="html-bibr">118</a>]. Copyright 2023, copy-right Xiaoxiao Xie. Adapted with permission from Ref. [<a href="#B119-nanomaterials-14-00473" class="html-bibr">119</a>]. Copyright 2022, copy-right Yifei Zhang. Adapted with permission from Ref. [<a href="#B83-nanomaterials-14-00473" class="html-bibr">83</a>]. Copyright 2023, copy-right Aaron Albert Aryee.</p>
Full article ">Figure 16
<p>Degradation efficiency of TC by different catalytic materials.</p>
Full article ">Figure 17
<p>(<b>a</b>) Degradation of AO7 by PMS with various catalysts. (<b>b</b>) Degradation of various organic dyes by FeNC-20/PMS [<a href="#B124-nanomaterials-14-00473" class="html-bibr">124</a>]. Adapted with permission from Ref. [<a href="#B124-nanomaterials-14-00473" class="html-bibr">124</a>]. Copyright 2022, copy-right Meng Li.</p>
Full article ">Figure 18
<p>(<b>a</b>) Different systems for the degradation of RhB. (<b>b</b>) Kinetic analysis [<a href="#B125-nanomaterials-14-00473" class="html-bibr">125</a>]. Adapted with permission from Ref. [<a href="#B125-nanomaterials-14-00473" class="html-bibr">125</a>]. Copyright 2023, copy-right Dengjie Zhong.</p>
Full article ">Figure 19
<p>(<b>a</b>,<b>b</b>) The degradation of RhB under different reaction conditions [<a href="#B127-nanomaterials-14-00473" class="html-bibr">127</a>]. Adapted with permission from Ref. [<a href="#B127-nanomaterials-14-00473" class="html-bibr">127</a>]. Copyright 2023, copy-right Ziyi Xiao.</p>
Full article ">Figure 20
<p>(<b>a</b>) Degradation of BPA with different catalytic conditions [<a href="#B6-nanomaterials-14-00473" class="html-bibr">6</a>]. (<b>b</b>) Removal efficiency of BPA in different systems [<a href="#B130-nanomaterials-14-00473" class="html-bibr">130</a>]. (<b>c</b>) Removal of TBBPA with different catalysts [<a href="#B131-nanomaterials-14-00473" class="html-bibr">131</a>]. Adapted with permission from Ref. [<a href="#B6-nanomaterials-14-00473" class="html-bibr">6</a>]. Copyright 2018, copy-right Yu Wang. Adapted with permission from Ref. [<a href="#B130-nanomaterials-14-00473" class="html-bibr">130</a>]. Copyright 2022, copy-right Yantao Wan. Adapted with permission from Ref. [<a href="#B131-nanomaterials-14-00473" class="html-bibr">131</a>]. Copyright 2020, copy-right Mei Huang.</p>
Full article ">Figure 21
<p>Degradation efficiency of SMX by different catalytic materials.</p>
Full article ">Figure 22
<p>(<b>a</b>) Effect of different radical quenchers on TC degradation in Mn-MIL-53(Fe)/PMS system. (<b>b</b>) EPR spectra obtained by using DMPO as spin-trapping agent [<a href="#B142-nanomaterials-14-00473" class="html-bibr">142</a>]. Adapted with permission from Ref. [<a href="#B142-nanomaterials-14-00473" class="html-bibr">142</a>]. Copyright 2022, copy-right Jun Yu.</p>
Full article ">Figure 23
<p>Possible reaction mechanism of TC degradation by Mn-MIL-53(Fe)/PMS system [<a href="#B142-nanomaterials-14-00473" class="html-bibr">142</a>].</p>
Full article ">Figure 24
<p>(<b>A</b>) Degradation of SMX in the presence of various quenching agents; (<b>B</b>) EPR spectra recorded in different activation systems (<b>a</b>) Fe@C-800/PS/MeOH/DMPO, (<b>b</b>) Fe@C-800/PS/DMPO, (<b>c</b>) Fe@C-800/PS/SMX/DMPO, (<b>d</b>) Fe@C-800/PS/TEMP, (<b>e</b>) Fe@C-800/PS/SMX/TEMP [<a href="#B21-nanomaterials-14-00473" class="html-bibr">21</a>]. Adapted with permission from Ref. [<a href="#B21-nanomaterials-14-00473" class="html-bibr">21</a>]. Copyright 2021, copy-right Mengjie Pu.</p>
Full article ">
17 pages, 2354 KiB  
Perspective
Research Progress on the Degradation of Organic Pollutants in Water by Activated Persulfate Using Biochar-Loaded Nano Zero-Valent Iron
by Hai Lu, Xiaoyan Wang, Qiao Cong, Xinglin Chen, Qingpo Li, Xueqi Li, Shuang Zhong, Huan Deng and Bojiao Yan
Molecules 2024, 29(5), 1130; https://doi.org/10.3390/molecules29051130 - 3 Mar 2024
Viewed by 1718
Abstract
Biochar (BC) is a new type of carbon material with a high specific surface area, porous structure, and good adsorption capacity, which can effectively adsorb and enrich organic pollutants. Meanwhile, nano zero-valent iron (nZVI) has excellent catalytic activity and can rapidly degrade organic [...] Read more.
Biochar (BC) is a new type of carbon material with a high specific surface area, porous structure, and good adsorption capacity, which can effectively adsorb and enrich organic pollutants. Meanwhile, nano zero-valent iron (nZVI) has excellent catalytic activity and can rapidly degrade organic pollutants through reduction and oxidation reactions. The combined utilization of BC and nZVI can not only give full play to their advantages in the adsorption and catalytic degradation of organic pollutants, but also help to reduce the agglomeration of nZVI, thus improving its efficiency in water treatment and providing strong technical support for water resources protection and environmental quality improvement. This article provides a detailed introduction to the preparation method and characterization technology, reaction mechanism, influencing factors, and specific applications of BC and nZVI, and elaborates on the research progress of BC-nZVI in activating persulfate (PS) to degrade organic pollutants in water. It has been proven experimentally that BC-nZVI can effectively remove phenols, dyes, pesticides, and other organic pollutants. Meanwhile, in response to the existing problems in current research, this article proposes future research directions and challenges, and summarizes the application prospects and development trends of BC-nZVI in water treatment. In summary, BC-nZVI-activated PS is an efficient technology for degrading organic pollutants in water, providing an effective solution for protecting water resources and improving environmental quality, and has significant application value. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Summary of major sources of biochar feedstock and its applications [<a href="#B21-molecules-29-01130" class="html-bibr">21</a>]. Reproduced with permission from [Yuan J, Wen Y, Dionysiou D D, Sharma V K, Ma X.], Chemical Engineering Journal]; published by [ELSEVIER], [2022].</p>
Full article ">Figure 2
<p>Mechanism of free radical pathway and non-free radical pathway in the nZVI/MoS2 BC+PS system [<a href="#B45-molecules-29-01130" class="html-bibr">45</a>]. Reproduced with permission from [Sun H, Zhang B, Wang N, Zhang N, Ma Y, Zang L, Li Z, Xue R], Journal of Environmental Chemical Engineering]; published by [ELSEVIER], [2023].</p>
Full article ">Figure 3
<p>Proposed mechanism of NP degradation by the nZVI/BC nanocomposite activation of persulfate [<a href="#B35-molecules-29-01130" class="html-bibr">35</a>]. Reproduced with permission from [Hussain I, Li M, Zhang Y, Li Y, Huang S, Du X, Liu G, Hayat W, Anwar N], Chemical Engineering Journal]; published by [ELSEVIER], [2017].</p>
Full article ">Figure 4
<p>Degradation rate of PNP by different systems [<a href="#B57-molecules-29-01130" class="html-bibr">57</a>]. Reproduced with permission from [Wang B, Zhu C, Ai D, Fan Z], Journal of Hazardous Materials]; published by [ELSEVIER], [2021].</p>
Full article ">Figure 5
<p>Degradation rate of AO II by different systems [<a href="#B59-molecules-29-01130" class="html-bibr">59</a>]. Reproduced with permission from [Liu H, Hu M, Zhang H, Wei J], Surfaces and Interfaces]; published by [ELSEVIER], [2022].</p>
Full article ">Figure 6
<p>Degradation rate of atrazine by different systems [<a href="#B61-molecules-29-01130" class="html-bibr">61</a>]. Reproduced with permission from [Jiang Z, Li J, Jiang D, Gao Y, Chen Y, Wang W, Cao B, Tao Y, Wang L, Zhang Y], Environmental Research]; published by [ELSEVIER], [2020].</p>
Full article ">Figure 7
<p>Degradation rate of SMZ by different systems [<a href="#B63-molecules-29-01130" class="html-bibr">63</a>]. Reproduced with permission from [Zhang L, Shen S], Journal of Industrial and Engineering Chemistry]; published by [ELSEVIER], [2020].</p>
Full article ">
14 pages, 5127 KiB  
Article
Preparation of Nickel-Based Bimetallic Catalyst and Its Activation of Persulfate for Degradation of Methyl Orange
by Bo Zhang, Jiale Li, Zhizhi Xu, Xiaohong Xu and Chundu Wu
Processes 2024, 12(2), 322; https://doi.org/10.3390/pr12020322 - 2 Feb 2024
Viewed by 1001
Abstract
In this research, a new catalyst for activating persulfate was developed by loading iron and nickel ions onto powdered activated carbon (PAC) for treating methyl orange, and the preparation process was optimized and characterized. The efficacy of the treatment was evaluated using the [...] Read more.
In this research, a new catalyst for activating persulfate was developed by loading iron and nickel ions onto powdered activated carbon (PAC) for treating methyl orange, and the preparation process was optimized and characterized. The efficacy of the treatment was evaluated using the Chemical Oxygen Demand (COD) removal rate, which reflects the impact of various process parameters, including catalyst dosage, sodium persulfate dosage, and reaction pH. Finally, the recovery and reuse performance of the catalyst were studied. The optimal conditions for preparing the activated sodium persulfate catalyst were determined to be as follows: a molar ratio of Fe3+ and Fe2+ to Ni of 4:1, a mass ratio of Fe3O4 to PAC of 1:4, a calcination temperature of 700 °C, and a calcination time of 4 h. This preparation led to an increase in surface porosity and the formation of a hollow structure within the catalyst. The active material on the surface was identified as nickel ferrite, comprising the elements C, O, Fe, and Ni. The magnetic property is beneficial to recycling. With the increase in catalyst and sodium persulfate dosage, the COD removal efficiency of the oxidation system increased first, and then, decreased. The catalyst showed good catalytic performance when the pH value was in the range of 3~11. Furthermore, Gas Chromatography–Mass Spectrometry (GC-MS) analysis indicated the complete oxidation of methyl orange dye molecules in the system. This result highlights the important role of the newly developed catalyst in activating persulfate. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of Ni doping amount on nickel-based bimetallic catalyst.</p>
Full article ">Figure 2
<p>Effect of PAC addition on nickel-based bimetallic catalyst.</p>
Full article ">Figure 3
<p>Effect of calcination temperature on nickel-based bimetallic catalyst.</p>
Full article ">Figure 4
<p>Effect of calcination time on nickel-based bimetallic catalyst.</p>
Full article ">Figure 5
<p>SEM pattern of PAC and nickel-based bimetallic catalyst: (<b>a</b>,<b>b</b>) nickel-based bimetallic catalyst; (<b>c</b>,<b>d</b>) PAC.</p>
Full article ">Figure 6
<p>TEM pattern of nickel-based bimetallic catalyst.</p>
Full article ">Figure 7
<p>EDS pattern of catalyst surface.</p>
Full article ">Figure 8
<p>XRD pattern of the catalyst.</p>
Full article ">Figure 9
<p>Effect of initial pH on COD removal rate.</p>
Full article ">Figure 10
<p>Effect of catalyst dosage on COD removal rate.</p>
Full article ">Figure 11
<p>Effect of sodium persulfate dosage on COD removal rate.</p>
Full article ">Figure 12
<p>Catalyst separation in magnetic field: (<b>a</b>) catalyst dispersed in solution; (<b>b</b>) catalyst separated in magnetic field.</p>
Full article ">Figure 13
<p>Effect of catalyst reuse on COD removal rate.</p>
Full article ">
13 pages, 1712 KiB  
Article
Development of Slow-Releasing Tablets Combined with Persulfate and Ferrous Iron for In Situ Chemical Oxidation in Trichloroethylene-Contaminated Aquifers
by Geumhee Yun, Sunhwa Park, Young Kim and Kyungjin Han
Water 2023, 15(23), 4103; https://doi.org/10.3390/w15234103 - 27 Nov 2023
Viewed by 1078
Abstract
Slow-releasing tablets combined with persulfate acting as an oxidant and ferrous iron acting as an activator were manufactured for in situ chemical oxidation. The trichloroethylene (TCE) removal efficiency according to the molar ratio of the oxidizer and activator in the 0, 0.5, 1, [...] Read more.
Slow-releasing tablets combined with persulfate acting as an oxidant and ferrous iron acting as an activator were manufactured for in situ chemical oxidation. The trichloroethylene (TCE) removal efficiency according to the molar ratio of the oxidizer and activator in the 0, 0.5, 1, 1.5, 2, and 2.5 molar ratio (persulfate: ferrous iron) reactors were 15%, 89%, 90%, 82%, 71%, and 55%, respectively. In a batch reactor injected with an oxidation-activation combined tablet (OACT) and a liquid oxidizing/activator, the TCE removal efficiencies were 100% and 70%, respectively, showing that the tablet form had a high efficiency in contaminant removal. The evaluation of the dissolution characteristics and TCE removal efficiency of OACT 0.5 (tablet with a 1:0.5 molar ratio of persulfate to activator) and OACT 1.0 (tablet with a 1:1 molar ratio of persulfate to activator) under continuous flow conditions showed that the TCE removal efficiency of the OACT 1.0 column was approximately 1.4 times higher than that of OACT 0.5. The longevities of persulfate and ferrous iron of the OACT 1.0 tablet were 43.2 days and 41.7 days, respectively. Thus, OACT 1.0, which was manufactured effectively, was suitable for in situ slow-release chemical oxidation systems. Full article
(This article belongs to the Topic Groundwater Pollution Control and Groundwater Management)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the column experiment.</p>
Full article ">Figure 2
<p>Release characteristics of prototypes SROT and SRAT under continuous flow conditions during the PTRT test (<b>a</b>) and TCE removal efficacy (<b>b</b>) and CO<sub>2</sub> concentrations (<b>c</b>) of prototype tablets in a batch reactor during the PTBT test.</p>
Full article ">Figure 3
<p>TCE degradation characteristics under various ferrous iron concentrations (<b>a</b>) and TCE degradation characteristics of prototype and combined tablets during the CTBT test (<b>b</b>).</p>
Full article ">Figure 4
<p>Normalized bromide and TCE concentrations in the soil column effluent during the CTST.</p>
Full article ">Figure 5
<p>Release characteristics of the combined tablet of (<b>a</b>) OACT 1.0 PS and (<b>b</b>) ferrous iron under continuous flow during CTST.</p>
Full article ">Figure 6
<p>Observed and predicted <span class="html-italic">M<sub>r</sub></span> values for persulfate and ferrous iron during continuous flow cell release experiments of CTST at a 0.1 mL/min flow rate.</p>
Full article ">
16 pages, 5225 KiB  
Article
Emerging Mesoporous Polyacrylamide/Gelatin–Iron Lanthanum Oxide Nanohybrids towards the Antibiotic Drugs Removal from the Wastewater
by Nazish Parveen, Fatimah Othman Alqahtani, Ghayah M. Alsulaim, Shada A. Alsharif, Kholoud M. Alnahdi, Hasna Abdullah Alali, Mohamad M. Ahmad and Sajid Ali Ansari
Nanomaterials 2023, 13(21), 2835; https://doi.org/10.3390/nano13212835 - 26 Oct 2023
Cited by 7 | Viewed by 1094
Abstract
The polyacrylamide/gelatin–iron lanthanum oxide (P-G-ILO nanohybrid) was fabricated by the free radical grafting co-polymerization technique in the presence of N,N-methylenebisacrylamide (MBA) as cross linker and ammonium persulfate (APS) as initiator. The P-G-ILO nanohybrid was characterized by the various spectroscopic and microscopic techniques that [...] Read more.
The polyacrylamide/gelatin–iron lanthanum oxide (P-G-ILO nanohybrid) was fabricated by the free radical grafting co-polymerization technique in the presence of N,N-methylenebisacrylamide (MBA) as cross linker and ammonium persulfate (APS) as initiator. The P-G-ILO nanohybrid was characterized by the various spectroscopic and microscopic techniques that provided the information regarding the crystalline behavior, surface area, and pore size. The response surface methodology was utilized for the statistical observation of diclofenac (DF) adsorption from the wastewater. The adsorption capacity (qe, mg/g) of P-G-ILO nanohybrid was higher (254, 256, and 258 mg/g) than the ILO nanoparticle (239, 234, and 233 mg/g). The Freundlich isotherm model was the best fitted, as it gives the higher values of correlation coefficient (R2 = 0.982, 0.991 and 0.981) and lower value of standard error of estimate (SEE = 6.30, 4.42 and 6.52), which suggested the multilayered adsorption of DF over the designed P-G-ILO nanohybrid and followed the pseudo second order kinetic model (PSO kinetic model) adsorption. The thermodynamic study reveals that adsorption was spontaneous and endothermic in nature and randomness onto the P-G-ILO nanohybrids surface increases after the DF adsorption. The mechanism of adsorption of DF demonstrated that the adsorption was mainly due to the electrostatic interaction, hydrogen bonding, and dipole interaction. P-G-ILO nanohybrid was reusable for up to five adsorption/desorption cycles. Full article
(This article belongs to the Section Environmental Nanoscience and Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD pattern of the ILO nanoparticles and P-G-ILO monohybrids and (<b>b</b>) FTIR spectra of the ILO nanoparticles, DF, P-G-ILO monohybrids, and DF-G-ILO monohybrids.</p>
Full article ">Figure 2
<p>The nitrogen adsorption/desorption graph for (<b>a</b>) ILO nanoparticles and (<b>b</b>) P-G-ILO nanohybrids, and (<b>c</b>) pore diameter profiles of the ILO nanoparticles and (<b>d</b>) P-G-ILO nanohybrids.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM image of the ILO nanoparticle, (<b>b</b>) SEM image of the P-G-ILO nanohybrid, (<b>c</b>) TEM image of the P-G-ILO nanohybrid, (<b>d</b>) EDX spectra of the ILO nanoparticle, and (<b>e</b>) EDX spectra of the P-G-ILO nanohybrid.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) Predicted vs. actual plot of <span class="html-italic">q</span><sub>e</sub> and 3D plots of the interaction of the independent factors and <span class="html-italic">q</span><sub>e</sub> at optimized conditions.</p>
Full article ">Figure 5
<p>(<b>a</b>) Langmuir isotherm, (<b>b</b>) Freundlich isotherm, (<b>c</b>) PFO kinetic model, and (<b>d</b>) PSO kinetic model plots of the DF adsorption over P-G-ILO nanohybrids.</p>
Full article ">Scheme 1
<p>Plausible mechanism of the adsorption of DF onto the P-G-ILO nanohybrid.</p>
Full article ">
33 pages, 15028 KiB  
Review
Recent Advances in Nanoscale Zero-Valent Iron (nZVI)-Based Advanced Oxidation Processes (AOPs): Applications, Mechanisms, and Future Prospects
by Mingyue Liu, Yuyuan Ye, Linli Xu, Ting Gao, Aiguo Zhong and Zhenjun Song
Nanomaterials 2023, 13(21), 2830; https://doi.org/10.3390/nano13212830 - 25 Oct 2023
Cited by 6 | Viewed by 2328
Abstract
The fast rise of organic pollution has posed severe health risks to human beings and toxic issues to ecosystems. Proper disposal toward these organic contaminants is significant to maintain a green and sustainable development. Among various techniques for environmental remediation, advanced oxidation processes [...] Read more.
The fast rise of organic pollution has posed severe health risks to human beings and toxic issues to ecosystems. Proper disposal toward these organic contaminants is significant to maintain a green and sustainable development. Among various techniques for environmental remediation, advanced oxidation processes (AOPs) can non-selectively oxidize and mineralize organic contaminants into CO2, H2O, and inorganic salts using free radicals that are generated from the activation of oxidants, such as persulfate, H2O2, O2, peracetic acid, periodate, percarbonate, etc., while the activation of oxidants using catalysts via Fenton-type reactions is crucial for the production of reactive oxygen species (ROS), i.e., •OH, •SO4, •O2, •O3CCH3, •O2CCH3, •IO3, •CO3, and 1O2. Nanoscale zero-valent iron (nZVI), with a core of Fe0 that performs a sustained activation effect in AOPs by gradually releasing ferrous ions, has been demonstrated as a cost-effective, high reactivity, easy recovery, easy recycling, and environmentally friendly heterogeneous catalyst of AOPs. The combination of nZVI and AOPs, providing an appropriate way for the complete degradation of organic pollutants via indiscriminate oxidation of ROS, is emerging as an important technique for environmental remediation and has received considerable attention in the last decade. The following review comprises a short survey of the most recent reports in the applications of nZVI participating AOPs, their mechanisms, and future prospects. It contains six sections, an introduction into the theme, applications of persulfate, hydrogen peroxide, oxygen, and other oxidants-based AOPs catalyzed with nZVI, and conclusions about the reported research with perspectives for future developments. Elucidation of the applications and mechanisms of nZVI-based AOPs with various oxidants may not only pave the way to more affordable AOP protocols, but may also promote exploration and fabrication of more effective and sustainable nZVI materials applicable in practical applications. Full article
(This article belongs to the Section Energy and Catalysis)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The diagram of the nZVI/MoS<sub>2</sub>-BC preparation; (<b>B</b>) TEM images of nZVI/MoS<sub>2</sub>-BC at 50 nm [<a href="#B84-nanomaterials-13-02830" class="html-bibr">84</a>], Copyright 2023 Elsevier. (<b>C</b>) SEM image of nZVI/Co<sub>3</sub>O<sub>4</sub> [<a href="#B60-nanomaterials-13-02830" class="html-bibr">60</a>], Copyright 2020 Elsevier. (<b>D</b>) TEM image of PVP-nZVI-Cu [<a href="#B71-nanomaterials-13-02830" class="html-bibr">71</a>], Copyright 2021 Elsevier. (<b>E</b>) SEM image of nZVI/CuO@BC, the insert comprises nZVI particles with diameter of ~50 nm that are uniformly dispersed on the surface and in the tubes of the CuO/BC composites [<a href="#B73-nanomaterials-13-02830" class="html-bibr">73</a>], Copyright 2021 Elsevier. (<b>F</b>) SEM image of nZVI@P-BC; (<b>G</b>) SEM image of nanocracks on nZVI nanosphere surface; (<b>H</b>) the formation mechanism of nanocracked nZVI [<a href="#B85-nanomaterials-13-02830" class="html-bibr">85</a>], Copyright 2023 Elsevier.</p>
Full article ">Figure 2
<p>(<b>A</b>) Effects of initial pH on TBBPA degradation efficiency [<a href="#B93-nanomaterials-13-02830" class="html-bibr">93</a>], Copyright 2021 Elsevier. (<b>B</b>) Influences of pH conditions on RhB degradation [<a href="#B84-nanomaterials-13-02830" class="html-bibr">84</a>], Copyright 2023 Elsevier. (<b>C</b>) Effects of solution pH on TCE removal in PVP-nZVI-Cu system [<a href="#B71-nanomaterials-13-02830" class="html-bibr">71</a>], Copyright 2021 Elsevier. (<b>D</b>) Degradation of atrazine with nZVI-Cu<sup>0</sup> at various temperatures [<a href="#B40-nanomaterials-13-02830" class="html-bibr">40</a>], Copyright 2023 Elsevier. (<b>E</b>) Effects of temperature on the degradation of norfloxacin [<a href="#B62-nanomaterials-13-02830" class="html-bibr">62</a>], Copyright 2020 Elsevier. (<b>F</b>) The Fe leaching of nZVI@NBC system and nZVI@BC system during the reaction under different initial pH values [<a href="#B77-nanomaterials-13-02830" class="html-bibr">77</a>], Copyright 2022 Elsevier.</p>
Full article ">Figure 3
<p>(<b>A</b>) Effects of catalyst dosage on TC catalytic degradation [<a href="#B60-nanomaterials-13-02830" class="html-bibr">60</a>], Copyright 2020 Elsevier. (<b>B</b>) Effects of PVP-nZVI-Cu dosage on TCE removal [<a href="#B71-nanomaterials-13-02830" class="html-bibr">71</a>], Copyright 2021 Elsevier. (<b>C</b>) Effects of different dosages of nZVI on degradation efficiency of TBBPA [<a href="#B93-nanomaterials-13-02830" class="html-bibr">93</a>], Copyright 2021 Elsevier. (<b>D</b>) Effects of different persulfate concentrations on TC catalytic degradation [<a href="#B60-nanomaterials-13-02830" class="html-bibr">60</a>], Copyright 2020 Elsevier. (<b>E</b>) Effects of different persulfate concentrations on TCE removal [<a href="#B71-nanomaterials-13-02830" class="html-bibr">71</a>], Copyright 2021 Elsevier. (<b>F</b>) Effects of different persulfate concentration on atrazine catalytic degradation using nZVI-Cu<sup>0</sup> [<a href="#B40-nanomaterials-13-02830" class="html-bibr">40</a>], Copyright 2023 Elsevier.</p>
Full article ">Figure 4
<p>(<b>A</b>) Effects of humic acid (HA) on the degradation of norfloxacin [<a href="#B62-nanomaterials-13-02830" class="html-bibr">62</a>], Copyright 2020 Elsevier. (<b>B</b>) Effects of natural organic matter (NOM) and (<b>C</b>) inorganic ions on TCE degradation in nZVI-Ni@BC-persulfate system [<a href="#B72-nanomaterials-13-02830" class="html-bibr">72</a>], Copyright 2021 Elsevier. (<b>D</b>) Levofloxacin degradation efficiency under different inorganic anions [<a href="#B69-nanomaterials-13-02830" class="html-bibr">69</a>], Copyright 2021 Elsevier. (<b>E</b>) Influencec of different anions on RhB degradation [<a href="#B84-nanomaterials-13-02830" class="html-bibr">84</a>], Copyright 2023 Elsevier. (<b>F</b>) Effects of co-existing anions and organic matter on BPA treatment performance with Fe<sup>0</sup>-CNTs catalytic membrane system [<a href="#B94-nanomaterials-13-02830" class="html-bibr">94</a>], Copyright 2022 Elsevier.</p>
Full article ">Figure 5
<p>(<b>A</b>) Reusability of nZVI-rGO under five consecutive oxidation cycles [<a href="#B58-nanomaterials-13-02830" class="html-bibr">58</a>], Copyright 2020 Elsevier. (<b>B</b>) Levofloxacin removal efficiency of nZVI/CF after 6 months (inset: the magnetic separation of the nZVI/CF dispersed in levofloxacin) [<a href="#B69-nanomaterials-13-02830" class="html-bibr">69</a>], Copyright 2021 Elsevier. (<b>C</b>) TCE removal in different cycles using PVP-ZVI-Cu nanoparticles in persulfate environment [<a href="#B71-nanomaterials-13-02830" class="html-bibr">71</a>], Copyright 2021 Elsevier. (<b>D</b>) Reusability study of nZVI-BC [<a href="#B87-nanomaterials-13-02830" class="html-bibr">87</a>], Copyright 2023 Elsevier. (<b>E</b>) Reusability of nZVI@NBC/persulfate system [<a href="#B77-nanomaterials-13-02830" class="html-bibr">77</a>], Copyright 2022 Elsevier. (<b>F</b>) Stability and durability of nZVI-Ni@BC catalyst [<a href="#B72-nanomaterials-13-02830" class="html-bibr">72</a>], Copyright 2021 Elsevier.</p>
Full article ">Figure 6
<p>(<b>A</b>) Proposed ROS evolution mechanisms in nZVI/yCo<sub>3</sub>O<sub>4</sub>/persulfate system [<a href="#B60-nanomaterials-13-02830" class="html-bibr">60</a>], Copyright 2020 Elsevier. (<b>B</b>) Degradation mechanisms of γ-HCH in the nZVI@P-BC/persulfate system [<a href="#B85-nanomaterials-13-02830" class="html-bibr">85</a>], Copyright 2023 Elsevier. (<b>C</b>) Mechanism scheme for phenol degradation in nZVI-BC/persulfate system [<a href="#B87-nanomaterials-13-02830" class="html-bibr">87</a>], Copyright 2023 Elsevier. (<b>D</b>) Proposed mechanisms of radical pathways and nonradical pathways in nZVI/MoS<sub>2</sub>-BC/persulfate system [<a href="#B84-nanomaterials-13-02830" class="html-bibr">84</a>], Copyright 2023 Elsevier.</p>
Full article ">Figure 7
<p>(<b>A</b>) Possible degradation pathway of BPA in nZVI@NBC-conducted AOP system [<a href="#B77-nanomaterials-13-02830" class="html-bibr">77</a>], Copyright 2022 Elsevier. (<b>B</b>) Degradation pathways of the RhB in nZVI/MoS<sub>2</sub>-conducted AOP system [<a href="#B84-nanomaterials-13-02830" class="html-bibr">84</a>], Copyright 2023 Elsevier.</p>
Full article ">Figure 8
<p>(<b>A</b>) Effects of different inorganic anions on the degradation of <span class="html-italic">p</span>-NP; (<b>B</b>) catalytic performance of rGO/PPy/nZVI catalytic microreactor in ultrapure water and East Lake water for long-term running [<a href="#B120-nanomaterials-13-02830" class="html-bibr">120</a>], Copyright 2023 Elsevier. Effects of (<b>C</b>) pH, (<b>D</b>) catalyst dosage, and (<b>E</b>) concentration of H<sub>2</sub>O<sub>2</sub> on the degradation of MB [<a href="#B116-nanomaterials-13-02830" class="html-bibr">116</a>], Copyright 2022 Elsevier. Effects of (<b>F</b>) temperature, (<b>G</b>) initial pH, (<b>H</b>) nZVI-BC dosage, and (<b>I</b>) H<sub>2</sub>O<sub>2</sub> concentration on the degradation of ornidazole [<a href="#B110-nanomaterials-13-02830" class="html-bibr">110</a>], Copyright 2020 MDPI.</p>
Full article ">Figure 9
<p>(<b>A</b>) Effects of PEG-nZVI@BC recycling on the degradation efficiency of 2,4-DCP; (<b>B</b>) the hysteresis loops of fresh and used PEG-nZVI@BC [<a href="#B121-nanomaterials-13-02830" class="html-bibr">121</a>], Copyright 2023 Springer. (<b>C</b>) Regeneration and reusability of nZVI-FBC nanocomposite [<a href="#B116-nanomaterials-13-02830" class="html-bibr">116</a>], Copyright 2022 Elsevier. (<b>D</b>) Recycling degradation of ornidazole with nZVI-BC [<a href="#B110-nanomaterials-13-02830" class="html-bibr">110</a>], Copyright 2020 MDPI. (<b>E</b>) Three cycles of degradation of SMX [<a href="#B129-nanomaterials-13-02830" class="html-bibr">129</a>], Copyright 2023 MDPI. (<b>F</b>) Long-term stability of rGO/PPy/nZVI and nZVI microreactor [<a href="#B120-nanomaterials-13-02830" class="html-bibr">120</a>], Copyright 2023 Elsevier.</p>
Full article ">Figure 10
<p>SEM image of Fe<sup>0</sup>-CNTs membrane: cross-section (<b>A</b>,<b>B</b>), outside surface (<b>C</b>); (<b>D</b>) mechanism of Fe<sup>0</sup>-CNTs catalytic membrane for BPA wastewater treatment; (<b>E</b>) removal of phenol, SMX, and paracetamol (ACM) with Fe<sup>0</sup>-CNTs catalytic membrane system [<a href="#B94-nanomaterials-13-02830" class="html-bibr">94</a>], Copyright 2022 Elsevier. SEM images of (<b>F</b>) upper surface, (<b>G</b>) cross-section, and (<b>H</b>,<b>I</b>) Fe NPs of nZVI-PVDF<sub>MW</sub> membrane; (<b>J</b>) scheme of cross-flow system employed for BPA removal experiments [<a href="#B128-nanomaterials-13-02830" class="html-bibr">128</a>], Copyright 2021 Elsevier.</p>
Full article ">Figure 11
<p>Scheme of the degradation of 1,2-DCA w nZVI/H<sub>2</sub>O<sub>2</sub> in source zones [<a href="#B115-nanomaterials-13-02830" class="html-bibr">115</a>], Copyright 2022 Elsevier.</p>
Full article ">Figure 12
<p>(<b>A</b>) Effects of dosage of 3D-GN@Fe/Al and (<b>B</b>) temperature on chloramphenicol removal [<a href="#B132-nanomaterials-13-02830" class="html-bibr">132</a>], Copyright 2021 Elsevier. (<b>C</b>) Effects of initial pH on catalytic reactivity of GT-nZVI/Cu [<a href="#B139-nanomaterials-13-02830" class="html-bibr">139</a>], Copyright 2020 Elsevier. (<b>D</b>) Effects of anions on MB decolorization efficiency ([salt]  =  0.1 M) and (<b>E</b>) effects of ionic strength on MB decolorization efficiency [<a href="#B137-nanomaterials-13-02830" class="html-bibr">137</a>], Copyright 2021 Springer. (<b>F</b>) Influences of co-existing ions and organic ligands on ciprofloxacin removal; (<b>G</b>) recycling of nZVI/PA composites [<a href="#B107-nanomaterials-13-02830" class="html-bibr">107</a>], Copyright 2023 MDPI. (<b>H</b>) Reusability of the Pt/nZVI composite [<a href="#B140-nanomaterials-13-02830" class="html-bibr">140</a>], Copyright 2020 Elsevier. (<b>I</b>) Effects of regeneration times on catalytic reactivity [<a href="#B139-nanomaterials-13-02830" class="html-bibr">139</a>], Copyright 2020 Elsevier.</p>
Full article ">Figure 13
<p>(<b>A</b>) Schematic model for removing ciprofloxacin by GT-nZVI/Cu under aerobic conditions [<a href="#B139-nanomaterials-13-02830" class="html-bibr">139</a>], Copyright 2020 Elsevier. (<b>B</b>) Reaction mechanism of A-nZVI and Sb(III) [<a href="#B138-nanomaterials-13-02830" class="html-bibr">138</a>], Copyright 2022 Elsevier. (<b>C</b>) Effects of DO on ciprofloxacin removal [<a href="#B107-nanomaterials-13-02830" class="html-bibr">107</a>], Copyright 2022 MDPI. (<b>D</b>) Contributions of adsorption, Fenton reaction, and reduction to overall OTC removal using Pt/nZVI composite [<a href="#B140-nanomaterials-13-02830" class="html-bibr">140</a>], Copyright 2020 Elsevier. (<b>E</b>) Possible degradation pathway of chloramphenicol with 3D-GN@Fe/Al [<a href="#B132-nanomaterials-13-02830" class="html-bibr">132</a>], Copyright 2021 MDPI.</p>
Full article ">Figure 14
<p>Effects of (<b>A</b>) solution pH, (<b>B</b>) nZVI dosage, and (<b>C</b>) PAA concentration on TC abatement in nZVI/PAA process; effects of (<b>D</b>) Cl<sup>−</sup>, (<b>E</b>) HPO<sub>4</sub><sup>2−</sup>, and (<b>F</b>) HA on TC abatement; (<b>G</b>) performance of reusable nZVI on TC abatement; (<b>H</b>) proposed mechanism for activation of PAA with Fe(II)-TC complexes in nZVI/PAA process [<a href="#B148-nanomaterials-13-02830" class="html-bibr">148</a>], Copyright 2022 Elsevier.</p>
Full article ">Figure 15
<p>Effects of (<b>A</b>) initial pH, (<b>B</b>) sodium percarbonate dosage, (<b>C</b>) TOCNF-Fe/Cu dosage, (<b>D</b>) Cl<sup>−</sup> concentration, (<b>E</b>) NO<sub>3</sub><sup>−</sup> concentration, (<b>F</b>) PO<sub>4</sub><sup>3−</sup> concentration, and (<b>G</b>) HA concentration on chloroform removal; (<b>H</b>) recyclability and (<b>I</b>) stability of TOCNF-Fe/Cu [<a href="#B151-nanomaterials-13-02830" class="html-bibr">151</a>], Copyright 2023 Elsevier.</p>
Full article ">Scheme 1
<p>Scheme of the oxidation of nZVI in solution and the structures of nZVI materials. (<b>A</b>) Polymer-stabilized nZVI; (<b>B</b>) metal-doped nZVI; (<b>C</b>) porous material-supported nZVI; (<b>D</b>) matrix-encapsulated nZVI; (<b>E</b>) chain-like structure of aggregated nZVI.</p>
Full article ">
17 pages, 3581 KiB  
Article
Efficient Degradation of Rhodamine B Dye through Hand Warmer Heterogeneous Activation of Persulfate
by Tiantian Ye, Lihong Liu, Yilin Wang, Jianqiang Zhang, Zhenxing Wang, Cong Li and Haoyu Luo
Sustainability 2023, 15(17), 13034; https://doi.org/10.3390/su151713034 - 29 Aug 2023
Cited by 2 | Viewed by 1083
Abstract
In this study, an innovative method for RhB (Rhodamine B) degradation in a persulfate (PS) and hand warmer heterogeneous activation system was investigated. The hand warmer showed better catalytic performance and excellent reusability in terms of PS activation. The reaction rate constants of [...] Read more.
In this study, an innovative method for RhB (Rhodamine B) degradation in a persulfate (PS) and hand warmer heterogeneous activation system was investigated. The hand warmer showed better catalytic performance and excellent reusability in terms of PS activation. The reaction rate constants of RhB removal in the hand warmer/PS process (0.354 min−1) were much faster than those in other PS- and Fe-related processes (0.010–0.233 min−1) at pH 7. The iron in the hand warmer is the main active ingredient to catalyze PS, and activated carbon, salt, and H+/OH accelerate the activated reaction due to the formation of micro-batteries in the solution. Moreover, the catalyst of the hand warmer showed excellent stability and reusability with a low level of iron leaching. This new, effective, inexpensive, repeatable, and environmentally friendly catalyst combined with PS has promising prospects for the removal of dyes from industrial wastewater. Full article
Show Figures

Figure 1

Figure 1
<p>SEM and corresponding EDS images of (<b>a</b>) the hand warmer and (<b>b</b>) oxidized hand warmer in ambient conditions.</p>
Full article ">Figure 2
<p>The XRD pattern of (<b>a</b>) the hand warmer and (<b>b</b>) oxidized hand warmer.</p>
Full article ">Figure 3
<p>XPS profiles of the surface of the hand warmer and oxidized hand warmer. (<b>a)</b> Full-range XPS spectra of hand warmer, (<b>b</b>) Fe 2p spectra of hand warmer, (<b>c</b>) C1s spectra of hand warmer, (<b>d</b>) Fe 2p spectra of oxidized hand warmer, (<b>e</b>) C1s spectra of oxidized hand warmer.</p>
Full article ">Figure 4
<p>FTIR spectra of hand warmer and oxidized hand warmer.</p>
Full article ">Figure 5
<p>Decolorization of RhB under different reaction systems (hand warmer, PS, PS/dark, PS/oxidized hand warmer, PS/hand warmer, PS/hand warmer/dark).</p>
Full article ">Figure 6
<p>Effects of several parameters on the decolorization of RhB under the PS/hand warmer system: (<b>a</b>) PS concentration, (<b>b</b>) the catalyst dosage, and (<b>c</b>) the initial pH. (<b>d</b>) Comparison of the apparent rate constants for the degradation of RhB in the different initial pH conditions.</p>
Full article ">Figure 7
<p>RhB removal in the presence of different radicals scavengers for the PS/hand warmer system. Conditions: RhB = 100 mg/L, PS = 5 mM, catalyst = 0.2 g/L, no pH adjustment.</p>
Full article ">Figure 8
<p>EPR spectra of DMPO-OH and DMPO-SO<sub>4</sub><sup>−</sup> adducts for various processes: (<b>a</b>) PS/hand warmer system, (<b>b</b>) PS/oxidized hand warmer system, (<b>c</b>) PS system.</p>
Full article ">Figure 9
<p>UV-vis spectral changes of RhB with PS/hand warmer system. Conditions: RhB = 100 mg/L, PS = 5 mM, catalyst = 0.2 g/L, no pH adjustment.</p>
Full article ">Figure 10
<p>(<b>a</b>) Mineralization of RhB in solution at different times; (<b>b</b>) the variation in pH in the PS/hand warmer.</p>
Full article ">Figure 11
<p>The reusability of the hand warmer. Reaction conditions: RhB 100 mg/L; PS 5 mM; hand warmer 0.4 g/L; no pH adjustment.</p>
Full article ">Scheme 1
<p>Possible activation mechanisms for radical generation in RhB degradation with (<b>a</b>) hand warmer in acidic conditions, (<b>b</b>) hand warmer in alkaline conditions, and (<b>c</b>) recycled hand warmer.</p>
Full article ">
15 pages, 9680 KiB  
Article
Degradation of 2-Chlorophenol in Aqueous Solutions Using Persulfate Activated by Biochar Supported Sulfide-Modified Nanoscale Zero-Valent Iron: Performance and Mechanisms
by Ronghuan Xie, Mu Wang, Weiping Li and Junjie Song
Water 2023, 15(15), 2805; https://doi.org/10.3390/w15152805 - 3 Aug 2023
Cited by 2 | Viewed by 1193
Abstract
In this work, soybean biochar-supported sulfide-modified nanoscale zero-valent iron (BC@S-nZVI) was synthesized and used to activate persulfate (PS) to degrade 2-chlorophenol (2-CP) in aqueous solutions. Batch experiments were carried out to investigate the degradation effects under different conditions, including initial mass ratios among [...] Read more.
In this work, soybean biochar-supported sulfide-modified nanoscale zero-valent iron (BC@S-nZVI) was synthesized and used to activate persulfate (PS) to degrade 2-chlorophenol (2-CP) in aqueous solutions. Batch experiments were carried out to investigate the degradation effects under different conditions, including initial mass ratios among 2-CP, PS, and BC@S-nZVI, initial pH values, temperature, and anions. The results showed that the mass ratio of PS to 2-CP equal to 70 and the mass ratio of BC@S-nZVI to PS equal to 0.4 were the optimum mass ratios in the degradation system. The degradation efficiency of 2-CP was higher under acidic and alkaline conditions than the neutral condition, and the effect was best at a pH of 3; meanwhile, it increased with the increase in temperature. Moreover, the degradation rate was restrained with the addition of Cl, promoted with the addition of NO3 and CO32−. Both free radical and material functions played leading roles in the degradation of 2-CP, and the stability of BC@S-nZVI was better than nZVI and S-nZVI. The experimental results showed that it was promising to remove 2-CP and other organic pollutants from groundwater by PS activated with BC@S-nZVI. Full article
(This article belongs to the Section Water Quality and Contamination)
Show Figures

Figure 1

Figure 1
<p>EDS and SEM images of BC (<b>a</b>,<b>d</b>), BC@S-nZVI before reaction (<b>b</b>,<b>e</b>), and BC@S-nZVI after reaction (<b>c</b>,<b>f</b>). The magnifications of the SEM images are all 10.00 k×.</p>
Full article ">Figure 2
<p>Effect of different activation systems on the degradation of 2-CP: (<b>a</b>) Different mass concentrations of PS, (<b>b</b>) different mass concentrations of BC@S-nZVI. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 2 Cont.
<p>Effect of different activation systems on the degradation of 2-CP: (<b>a</b>) Different mass concentrations of PS, (<b>b</b>) different mass concentrations of BC@S-nZVI. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 3
<p>Degradation effects of various reaction systems. Reaction conditions: The mass ratio of PS to 2-CP was 70 and the mass ratio of BC@S-nZVI to PS was 0.4, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 4
<p>Degradation effects of different initial concentration values of 2-CP. Reaction conditions: The mass ratio of PS to 2-CP was 70 and the mass ratio of BC@S-nZVI to PS was 0.4, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 5
<p>Effect of pH and temperature on the degradation of 2-CP: (<b>a</b>) Different values of pH. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, T = 30 °C; (<b>b</b>) different values of temperature. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, pH = 5 ± 0.2.</p>
Full article ">Figure 6
<p>Effect of inorganic anions on the degradation of 2-CP (<b>a</b>) Cl<sup>−</sup>, (<b>b</b>) NO<sub>3</sub><sup>−</sup>, and (<b>c</b>) CO<sub>3</sub><sup>2−</sup>. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 6 Cont.
<p>Effect of inorganic anions on the degradation of 2-CP (<b>a</b>) Cl<sup>−</sup>, (<b>b</b>) NO<sub>3</sub><sup>−</sup>, and (<b>c</b>) CO<sub>3</sub><sup>2−</sup>. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 7
<p>ESR spectra of DMPO-HO• and DMPO-SO4•<sup>−</sup> adduct. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, DMPO = 2800 mg/L, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 8
<p>The removal rates of 2-CP in different reaction systems. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 9
<p>Contribution rate of chemical factors at different reaction times. Reaction conditions: [2-CP] = 20 mg/L of 20 mL, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">Figure 10
<p>Reusability of catalysts in three cycles: (<b>a</b>) nZVI/PS, (<b>b</b>) S-nZVI/PS, and (<b>c</b>) BC@S-nZVI/PS. Reaction conditions: [2-CP] = 20 mg/L, PS = 1400 mg/L, BC@S-nZVI = 560 mg/L, T = 30 °C, pH = 5 ± 0.2.</p>
Full article ">
14 pages, 5786 KiB  
Article
The Sonocatalytic Activation of Persulfates on Iron Nanoparticle Decorated Zeolite for the Degradation of 1,4-Dioxane in Aquatic Environments
by Surya Teja Malkapuram, Shirish Hari Sonawane, Manoj P. Rayaroth, Murali Mohan Seepana, Sivakumar Manickam, Jakub Karczewski and Grzegorz Boczkaj
Catalysts 2023, 13(7), 1065; https://doi.org/10.3390/catal13071065 - 1 Jul 2023
Cited by 3 | Viewed by 1593
Abstract
In the chemical industry, 1,4-diethylene dioxide, commonly called dioxane, is widely used as a solvent as well as a stabilizing agent for chlorinated solvents. Due to its high miscibility, dioxane is a ubiquitous water contaminant. This study investigates the effectiveness of catalyst- and [...] Read more.
In the chemical industry, 1,4-diethylene dioxide, commonly called dioxane, is widely used as a solvent as well as a stabilizing agent for chlorinated solvents. Due to its high miscibility, dioxane is a ubiquitous water contaminant. This study investigates the effectiveness of catalyst- and ultrasound (US)-assisted persulfate (PS) activation with regard to degrading dioxane. As a first step, a composite catalyst was prepared using zeolite. A sonochemical dispersion and reduction method was used to dope zeolite with iron nanoparticles (FeNP/Z). In the subsequent study, the reaction kinetics of dioxane degradation following the single-stage and two-stage addition of PS was examined in the presence of a catalyst. Using GC-MS analysis, intermediate compounds formed from dioxane degradation were identified, and plausible reaction pathways were described. Upon 120 min of sonication in the presence of a catalyst with a two-stage injection of PS, 95% 100 mg/L dioxane was degraded. Finally, the estimated cost of treatment is also reported in this study. Sonolytically activated PS combined with a FeNP/Z catalyst synergizes the remediation of biorefractory micropollutants such as dioxane. Full article
(This article belongs to the Section Environmental Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of iron impregnated molecular sieve.</p>
Full article ">Figure 2
<p>SEM analysis of the surface morphology of FeNP/Z.</p>
Full article ">Figure 3
<p>Elemental analysis of FeNP/Z using EDS.</p>
Full article ">Figure 4
<p>XPS spectrum of synthesized composite catalyst for the identification of oxidized forms of iron.</p>
Full article ">Figure 5
<p>Degradation of dioxane using US, US + PS, and US + PS + Cat.</p>
Full article ">Figure 6
<p>Degradation kinetics of dioxane (K<sub>US</sub> &lt; K<sub>US+PS</sub> &lt; K<sub>US+PS+Cat</sub>).</p>
Full article ">Figure 7
<p>Degradation of dioxane in single and double stages.</p>
Full article ">Figure 8
<p>A possible pathway for the degradation of dioxane in the presence of a catalyst.</p>
Full article ">Figure 9
<p>Schematic diagram for FeNP/Z synthesis.</p>
Full article ">Figure 10
<p>Ultrasonic experimental setup for the degradation of dioxane.</p>
Full article ">
16 pages, 3219 KiB  
Article
Efficient Remediation of p-chloroaniline Contaminated Soil by Activated Persulfate Using Ball Milling Nanosized Zero Valent Iron/Biochar Composite: Performance and Mechanisms
by Zihan Guo, Dong Wang, Zichen Yan, Linbo Qian, Lei Yang, Jingchun Yan and Mengfang Chen
Nanomaterials 2023, 13(9), 1517; https://doi.org/10.3390/nano13091517 - 29 Apr 2023
Cited by 2 | Viewed by 1590
Abstract
In this study, efficient remediation of p-chloroaniline (PCA)-contaminated soil by activated persulfate (PS) using nanosized zero-valent iron/biochar (B-nZVI/BC) through the ball milling method was conducted. Under the conditions of 4.8 g kg−1 B-nZVI/BC and 42.0 mmol L−1 PS with pH [...] Read more.
In this study, efficient remediation of p-chloroaniline (PCA)-contaminated soil by activated persulfate (PS) using nanosized zero-valent iron/biochar (B-nZVI/BC) through the ball milling method was conducted. Under the conditions of 4.8 g kg−1 B-nZVI/BC and 42.0 mmol L−1 PS with pH 7.49, the concentration of PCA in soil was dramatically decreased from 3.64 mg kg−1 to 1.33 mg kg−1, which was much lower than the remediation target value of 1.96 mg kg−1. Further increasing B-nZVI/BC dosage and PS concentration to 14.4 g kg−1 and 126.0 mmol L−1, the concentration of PCA was as low as 0.15 mg kg−1, corresponding to a degradation efficiency of 95.9%. Electron paramagnetic resonance (EPR) signals indicated SO4, •OH, and O2 radicals were generated and accounted for PCA degradation with the effect of low-valence iron and through the electron transfer process of the sp2 hybridized carbon structure of biochar. 1-chlorobutane and glycine were formed and subsequently decomposed into butanol, butyric acid, ethylene glycol, and glycolic acid, and the degradation pathway of PCA in the B-nZVI/BC-PS system was proposed accordingly. The findings provide a significant implication for cost-effective and environmentally friendly remediation of PCA-contaminated soil using a facile ball milling preparation of B-nZVI/BC and PS. Full article
(This article belongs to the Section Environmental Nanoscience and Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM image (<b>a</b>), TEM image (<b>b</b>), and TEM-EDS (<b>c</b>–<b>f</b>) images of B-nZVI/BC.</p>
Full article ">Figure 2
<p>XRD patterns (<b>a</b>), FT-IR spectra (<b>b</b>), Raman spectra (<b>c</b>), and BET nitrogen adsorption-desorption isotherms (<b>d</b>) of ZVI, BC, and B-nZVI/BC.</p>
Full article ">Figure 2 Cont.
<p>XRD patterns (<b>a</b>), FT-IR spectra (<b>b</b>), Raman spectra (<b>c</b>), and BET nitrogen adsorption-desorption isotherms (<b>d</b>) of ZVI, BC, and B-nZVI/BC.</p>
Full article ">Figure 3
<p>Kinetics curves of the PCA removal (<b>a</b>) and pseudo-first-order kinetics of PCA in ZVI, BC, and B-nZVI/BC activated PS systems (<b>b</b>). Reaction conditions: [PS]<sub>0</sub> = 42.0 mmol L<sup>−1</sup>, [B-nZVI/BC]<sub>0</sub> = 4.8 g kg<sup>−1</sup>, [ZVI]<sub>0</sub> = 2.4 g kg<sup>−1</sup>, [BC]<sub>0</sub> = 2.4 g kg<sup>−1</sup>, pH<sub>0</sub> = 7.49, and T = 25 °C.</p>
Full article ">Figure 4
<p>Effects of B-nZVI/BC dosage (<b>a</b>) and PS concentration (<b>b</b>) on the degradation of PCA in the B-nZVI/BC-PS system. Reaction conditions: [PS]<sub>0</sub> = 42.0 mmol L<sup>−1</sup> for (<b>a</b>), [B-nZVI/BC]<sub>0</sub> = 4.8 g kg<sup>−1</sup> for (<b>b</b>), pH<sub>0</sub> = 7.49, and T = 25 °C.</p>
Full article ">Figure 5
<p>XPS survey spectra (<b>a</b>), Fe 2p (<b>b</b>), C 1s (<b>c</b>), and O 1s (<b>d</b>) spectra of B-nZVI/BC before and after reaction.</p>
Full article ">Figure 6
<p>EPR measurements for SO<sub>4</sub>•<sup>−</sup>, •OH (<b>a</b>), and O<sub>2</sub>•<sup>−</sup> (the EPR spectrometric detection of O<sub>2</sub>•<sup>−</sup> was performed in DMSO solution, which was designed to avoid the influence of SO<sub>4</sub>•<sup>−</sup> and •OH) (<b>b</b>) in various activated PS systems. Reaction conditions: [PS]<sub>0</sub> = 42.0 mmol L<sup>−1</sup>, [DMSO]: [H<sub>2</sub>O] =9:1 (volume ratio), [DMPO] = 200.0 mmol L<sup>−1</sup>, [B-nZVI/BC]<sub>0</sub> = 4.8 g kg<sup>−1</sup>, pH<sub>0</sub> = 7.49, and T = 25 °C.</p>
Full article ">Figure 7
<p>Degradation pathways of PCA in the B-nZVI/BC-PS system.</p>
Full article ">
10 pages, 2215 KiB  
Article
Magnetic Nanocomposites for the Remote Activation of Sulfate Radicals for the Removal of Rhodamine B
by Pranto Paul, Marissa Nicholson and J. Zach Hilt
Nanomaterials 2023, 13(7), 1151; https://doi.org/10.3390/nano13071151 - 23 Mar 2023
Cited by 1 | Viewed by 1456
Abstract
The widespread presence of numerous organic contaminants in water poses a threat to the ecological environment and human health. Magnetic nanocomposites exposed to an alternating magnetic field (AMF) have a unique ability for magnetically mediated energy delivery (MagMED) resulting from the embedded magnetic [...] Read more.
The widespread presence of numerous organic contaminants in water poses a threat to the ecological environment and human health. Magnetic nanocomposites exposed to an alternating magnetic field (AMF) have a unique ability for magnetically mediated energy delivery (MagMED) resulting from the embedded magnetic nanoparticles; this localized energy delivery and associated chemical and thermal effects are a potential method for removing contaminants from water. This work developed a novel magnetic nanocomposite—a polyacrylamide-based hydrogel loaded with iron oxide nanoparticles. For this magnetic nanocomposite, persulfate activation and the contamination removal in water were investigated. Magnetic nanocomposites were exposed to AMF with a model organic contaminant, rhodamine B (RhB) dye, with or without sodium persulfate (SPS). The removal of RhB by the nanocomposite without SPS as a sorbent was found to be proportional to the concentration of magnetic nanoparticles (MNPs) in the nanocomposite. With the addition of SPS, approximately 100% of RhB was removed within 20 min. This removal was attributed primarily to the activation of sulfate radicals, triggered by MNPs, and the localized heating resulted from the MNPs when exposed to AMF. This suggests that this magnetic nanocomposite and an AMF could be a unique environmental remediation technique for hazardous contaminants. Full article
(This article belongs to the Special Issue Polymer Based Nanocomposites: Experiment, Theory and Simulations)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the AMF treatment with magnetic nanocomposite and sodium persulfate to remove RhB. Created using biorender.com.</p>
Full article ">Figure 2
<p>Kinetic swelling study of IONP-acrylamide nanocomposite exposed to different temperatures. (N = 3, error bars represent the standard deviation).</p>
Full article ">Figure 3
<p>Removal of RhB without SPS (<b>a</b>) at different MNP concentrations in the nanocomposites in an AMF strength of 34.13 kA/m; (<b>b</b>) at different strengths of AMF with a nanocomposite of 3% MNP. (N = 3, error bars represent the standard deviation).</p>
Full article ">Figure 4
<p>Removal of RhB with SPS (<b>a</b>) at different strengths of AMF with a nanocomposite of 3% MNP; (<b>b</b>) with different MNP concentrations in nanocomposites with an AMF strength of 34.13 kA/m. Pseudo-first-order plot of the RhB degradation (<b>c</b>) at different strengths of AMF with a nanocomposite of 3% MNP; (<b>d</b>) at different MNP concentrations in the nanocomposite with an AMF strength of 34.13 kA/m; (N = 3, error bars represent the standard deviation).</p>
Full article ">Figure 5
<p>Potential mechanism for the production of enhanced sulfate radicals for RhB degradation. Created using biorender.com.</p>
Full article ">
Back to TopTop