Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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 (217)

Search Parameters:
Keywords = peroxymonosulfate

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 5879 KiB  
Article
Comparison of UV/PAA and VUV/PAA Processes for Eliminating Diethyl Phthalate in Water
by Feilong Dong, Jiayi Cheng, Yifeng Cheng and Xiaoyan Ma
Water 2024, 16(23), 3533; https://doi.org/10.3390/w16233533 - 8 Dec 2024
Viewed by 396
Abstract
Diethyl phthalate (DEP) is a commonly utilized plasticizer that has gained significant attention due to its widespread occurrence in the environment and its harmful impact on human health. The primary objective of this study was to evaluate and compare several (ultraviolet) UV-(peracetic acid) [...] Read more.
Diethyl phthalate (DEP) is a commonly utilized plasticizer that has gained significant attention due to its widespread occurrence in the environment and its harmful impact on human health. The primary objective of this study was to evaluate and compare several (ultraviolet) UV-(peracetic acid) PAA advanced oxidation processes based on hydroxyl radicals to degrade DEP. The effect of UV-LEDs incorporating PAA at different UV ranges (UV-A, λ = 365 nm; UV-C, λ = 254 nm and VUV, λ = 254 nm) was evaluated. The results demonstrated that DEP was successfully degraded in both the UVC/PAA (removal rate 98.28%) and VUV/PAA (removal rate 97.72%) processes compared to the UVA/PAA process (removal rate of 2.71%). The competitive method evaluated the contribution of R-O•, which were 24.08% and 33.92% in UVC/PAA and VUV/PAA processes, respectively. We also evaluated the effects of peroxymonosulfate (PMS) dosages, UV irradiation, pH and anion coexistence on the removal of DEP. In the UVC/PAA system, DEP degradation was particularly effective (removal rate about 95.52%) over a wider pH range (3–9). As the concentration of HCO3 ions increased, there may have been some inhibition of DEP removal. The inhibitory effect of HA and Cl ions on DEP removal were negligible. Analysis of the intermediates revealed that DEP degradation primarily occurred via two pathways: hydrolysis and hydroxylation reactions. This study presents a potential mnethod for the removal of phthalates and offers some guidance for the selection of appropriate disinfection technologies in drinking water treatment. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Degradation of DEP in the UVA, UVC, VUV, PAA, UVA/PAA, UVC/PAA, and VUV/PAA systems (<b>b</b>) corresponding kinetic analysis in the UVA, UVC, VUV, PAA, UVA/PAA, UVC/PAA, and VUV/PAA systems (conditions: [DEP]<sub>0</sub> = 2.0 μM, [PAA]<sub>0</sub> = 0.5 mM and pH = 9.0).</p>
Full article ">Figure 2
<p>The competitive degradations of pCBA and DEP in the UVC/PAA (<b>a</b>) and VUV/PAA (<b>b</b>) systems, the degradation rates of pCBA and DEP in the (<b>c</b>) UVC/PAA and (<b>d</b>) VUV/PAA systems. (conditions: [DEP]<sub>0</sub> = 2.0 μM, [PAA]<sub>0</sub> = 0.5 mM, [TBA] = 10 mM, [MeOH] = 10 mM and pH = 9.0).</p>
Full article ">Figure 3
<p>Degradation of DEP (<b>a</b>) and corresponding kinetic analysis (<b>b</b>) under different PAA concentrations (0.125–1.0 mM) in UVC/PAA system; Degradation of DEP (<b>c</b>) and corresponding kinetic analysis (<b>d</b>) under different UVC light intensity (378, 615 and 855 μW/cm<sup>2</sup>) in UVC/PAA system; Degradation of DEP (<b>e</b>) and corresponding kinetic analysis (<b>f</b>) under different pH (3.0, 5.0, 7.0, 9.0 and 11.0) in UVC/PAA system (conditions: [DEP]<sub>0</sub> = 2.0 μM, [PAA]<sub>0</sub> = 0.5 mM, and pH = 9.0).</p>
Full article ">Figure 4
<p>Proposed reaction pathways of DEP degradation in the UVC/PAA system.</p>
Full article ">Figure 5
<p>Degradation of DEP under different concentrations of Cl<sup>−</sup> (<b>a</b>), HCO<sub>3</sub><sup>−</sup> (<b>b</b>), HNO<sub>3</sub><sup>−</sup> (<b>c</b>) and HA (<b>d</b>) in UVC/PAA process (conditions: [DEP]<sub>0</sub> = 2.0 μM and [PAA]<sub>0</sub> = 0.5 mM).</p>
Full article ">Figure 6
<p>Energy consumption evaluation during DEP degradation in the UVC/PAA and VUV/PAA systems.</p>
Full article ">
21 pages, 4254 KiB  
Article
Effects of Manganese Carbonate Addition on the Carbocatalytic Properties of Lignocellulosic Waste for Use in the Degradation of Acetaminophen
by Camila Mosquera-Olano, Carolina Quimbaya, Vanessa Rodríguez, Angie Vanessa-Lasso, Santiago Correa, E. D. C. Castrillón, John Rojas and Yenny P. Ávila-Torres
Polymers 2024, 16(23), 3316; https://doi.org/10.3390/polym16233316 - 27 Nov 2024
Viewed by 402
Abstract
A carbon-based material was synthesized using potato peels (BPP) and banana pseudo-stems (BPS), both of which were modified with manganese to produce BPP-Mn and BPS-Mn, respectively. These materials were assessed for their ability to activate peroxymonosulfate (PMS) in the presence of MnCO3 [...] Read more.
A carbon-based material was synthesized using potato peels (BPP) and banana pseudo-stems (BPS), both of which were modified with manganese to produce BPP-Mn and BPS-Mn, respectively. These materials were assessed for their ability to activate peroxymonosulfate (PMS) in the presence of MnCO3 to degrade acetaminophen (ACE), an emerging water contaminant. The materials underwent characterization using spectroscopic, textural, and electrochemical techniques. Manganese served a dual function: enhancing adsorption properties and facilitating the breaking of peroxide bonds. Additionally, carbonate ions played a structural role in the materials, transforming into CO2 at high temperatures and thereby increasing material porosity, which improved adsorption capabilities. This presents a notable advantage for materials that have not undergone de-lignification. Among the materials tested, BPS exhibited the highest efficiency in the carbocatalytic degradation of ACE, achieving a synergy index of 1.31 within just 5 min, with 42% ACE degradation in BPS compared to BPS-Mn, which achieved 100% ACE removal through adsorption. Reactive oxygen species such as sulfate, hydroxyl, and superoxide anion radicals were identified as the primary contributors to pollutant degradation. In contrast, no degradation was observed for BPP and BPP-Mn, which is likely linked to the lower lignin content in their precursor material. This work addressed the challenge of revalorizing lignocellulosic waste by highlighting its potential as an oxidant for emerging pollutants. Furthermore, the study demonstrated the coexistence of various reactive oxygen species, confirming the capacity of carbon-based matrices to activate PMS. Full article
(This article belongs to the Special Issue Lignin: Modifications and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>). PP and (<b>b</b>). PP without lignin using alkali treatment.</p>
Full article ">Figure 2
<p>BET adsorption-desorption and size pore distribution for (<b>a</b>). BPP, (<b>b</b>). BPP-Mn, (<b>c</b>). BPS, (<b>d</b>). BPS-Mn.</p>
Full article ">Figure 2 Cont.
<p>BET adsorption-desorption and size pore distribution for (<b>a</b>). BPP, (<b>b</b>). BPP-Mn, (<b>c</b>). BPS, (<b>d</b>). BPS-Mn.</p>
Full article ">Figure 3
<p>SEM- images of raw materials, carbonaceous materials, exfoliated and modified with manganese. (<b>a</b>). PP, (<b>b</b>). BPP, (<b>c</b>). BPP-Mn, (<b>d</b>) PS, (<b>e</b>). BPS, (<b>f</b>). BPS-Mn.</p>
Full article ">Figure 4
<p>XRD patterns for (<b>a</b>). BPS-Mn, (<b>b</b>). BPS, (<b>c</b>). BPP-Mn, (<b>d</b>). BPP. The green dotted line represents Graphitic phase.</p>
Full article ">Figure 5
<p>Remotion of ACE with carbonaceous material, (<b>a</b>). BPP and BPP-Mn, (<b>b</b>). BPS and BPS-Mn.</p>
Full article ">Figure 6
<p>Voltammetry cycles for material activation and carbocatalytic process with and without PMS, scanning rate of 20 mV s<sup>−1</sup>, −2.0 V, and +2.0 V as support electrolyte: [Na<sub>2</sub>SO<sub>4</sub>] = 0.1 M, (<b>a</b>). BPP + PMS, (<b>b</b>). BPP + PMS + ACE, (<b>c</b>). BPP<sup>_</sup> Mn+ PMS, (<b>d</b>). BPP<sup>_</sup> Mn + PMS + ACE, (<b>e</b>). BPS + PMS, (<b>f</b>). BPS + PMS + ACE, (<b>g</b>). BPS<sup>_</sup> Mn + PMS + ACE and (<b>h</b>). BPS <sup>_</sup>Mn + PMS + ACE.</p>
Full article ">Figure 6 Cont.
<p>Voltammetry cycles for material activation and carbocatalytic process with and without PMS, scanning rate of 20 mV s<sup>−1</sup>, −2.0 V, and +2.0 V as support electrolyte: [Na<sub>2</sub>SO<sub>4</sub>] = 0.1 M, (<b>a</b>). BPP + PMS, (<b>b</b>). BPP + PMS + ACE, (<b>c</b>). BPP<sup>_</sup> Mn+ PMS, (<b>d</b>). BPP<sup>_</sup> Mn + PMS + ACE, (<b>e</b>). BPS + PMS, (<b>f</b>). BPS + PMS + ACE, (<b>g</b>). BPS<sup>_</sup> Mn + PMS + ACE and (<b>h</b>). BPS <sup>_</sup>Mn + PMS + ACE.</p>
Full article ">Figure 7
<p>Electronic spectra corresponding to the interaction between ACE-PMS-BPS-Mn for t = 0 min–2 min.</p>
Full article ">Figure 8
<p>Mechanism elucidation for BPS material (<b>a</b>) using scavengers for hydroxyl radical, sulfate radical, and singlet oxygen, (<b>b</b>) Electrochemical impedance for BPS, (<b>c</b>) and RAMAN shift for BPS after/before treatment. The blue dotted line represents Graphitic phase. The red dotted line represents Graphitic phaseGraphitic structure Diamond structure.</p>
Full article ">Scheme 1
<p>Mechanism proposed BPS and its modification with manganese.</p>
Full article ">
13 pages, 3601 KiB  
Article
Rapid Degradation of Bisphenol F Using Magnetically Separable Bimetallic Biochar Composite Activated by Peroxymonosulfate
by Hui Liang, Ruijuan Li, Tongjin Liu, Rumei Li, Yuxiao Zhu and Feng Fang
Molecules 2024, 29(23), 5545; https://doi.org/10.3390/molecules29235545 - 24 Nov 2024
Viewed by 540
Abstract
Peroxymonosulfate (PMS)-based advanced oxidation processes have shown potential for the removal of organic contaminants; however, the preparation of catalysts with high degradation efficiencies and rapid reaction rates remains a challenge. In this study, we have successfully synthesized CoFe bimetallic modified corn cob-derived biochar [...] Read more.
Peroxymonosulfate (PMS)-based advanced oxidation processes have shown potential for the removal of organic contaminants; however, the preparation of catalysts with high degradation efficiencies and rapid reaction rates remains a challenge. In this study, we have successfully synthesized CoFe bimetallic modified corn cob-derived biochar (CoFe/BC) for the activation of PMS, achieving the rapid and efficient degradation of bisphenol F (BPF). The synthesized CoFe/BC catalyst demonstrated excellent catalytic performance, achieving over 99% removal within 3 min and exhibiting a removal rate of 90.0% after five cycles. This could be attributed to the cyclic transformation of Co and Fe, which sustained rapid PMS activation for BPF degradation, and Co0/Fe0 played a significant role in the cyclic transformation. Furthermore, the electron paramagnetic resonance tests confirmed that •SO4 and •OH were the primary reactive oxygen species, while •O2 played a minor role in BPF degradation. This study highlights the high degradation efficiency, rapid reaction rate, excellent magnetic separation properties, and exceptional reusability of CoFe/BC catalysts for BPF removal, providing valuable insights for practical wastewater treatment. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>d</b>) SEM images, (<b>e</b>–<b>i</b>) EDS mappings (C, N, O, Co, Fe), and (<b>j</b>–<b>l</b>) HRTEM images of CoFe/BC catalyst.</p>
Full article ">Figure 2
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isothermal curve and (<b>b</b>) pore-size distribution curve of CoFe/BC catalyst; (<b>c</b>) XRD patterns of BC, CoFe<sub>2</sub>O<sub>4</sub>/CoFe/BC (Co:Fe = 1:9), CoFe/BC37 (Co:Fe = 3:7), CoFe/BC (Co:Fe = 5:5), CoFe/BC73 (Co:Fe = 7:3), and CoC<sub>x</sub>/Co<sub>7</sub>Fe<sub>3</sub>/BC (Co:Fe = 9:1) catalysts. The proportion of Co and Fe mentioned in this study is the molar ratio, and further information on the catalysts is available in <a href="#sec3dot2-molecules-29-05545" class="html-sec">Section 3.2</a>. Synthesis.</p>
Full article ">Figure 3
<p>(<b>a</b>) XPS survey, (<b>b</b>) C 1s, (<b>c</b>) O 1s, (<b>d</b>) N 1s, (<b>e</b>) Co 2p, and (<b>f</b>) Fe 2p of CoFe/BC catalyst.</p>
Full article ">Figure 4
<p>(<b>a</b>) Degradation efficiencies of BPF degradation on different catalysts. Influence of various conditions on BPF degradation in CoFe/BC/PMS system: (<b>b</b>) catalyst amount, (<b>c</b>) PMS concentration, and (<b>d</b>) reaction temperature. Experimental conditions: Catalyst = 0.05 g, BPF = 10 mg/L, PMS = 100 mg/L, T = 25 °C.</p>
Full article ">Figure 5
<p>EPR spectra of (<b>a</b>) DMPO-X, (<b>b</b>) DMPO-•O<sub>2</sub><sup>−</sup>, and (<b>c</b>) TEMP-<sup>1</sup>O<sub>2</sub>; (<b>d</b>) the proposed mechanism for BPF degradation over the CoFe/BC/PMS system.</p>
Full article ">Figure 6
<p>(<b>a</b>) Recyclability of CoFe/BC for BPF removal and (<b>b</b>) corresponding k value; (<b>c</b>) XRD spectrum of CoFe/BC after reaction; (<b>d</b>) degradation performance of CoFe/BC catalyst on different organic pollutants. Experimental conditions: Catalyst = 0.05 g, BPF = 10 mg/L, BPA = 10 mg/L, phenol = 10 mg/L, MB = 50 mg/L, MG = 50 mg/L, RhB = 50 mg/L, PMS = 100 mg/L, T = 25 °C.</p>
Full article ">
14 pages, 4643 KiB  
Article
Degradation of Tetracycline (TC) by ZrO2-3DG/PMS System: Revealing the Role of Defects in the Conditions of Light Irradiation and Sulfate Accumulation
by Jixiang Duan, Xin Wang, Zhihong Ye and Fuming Chen
Catalysts 2024, 14(12), 846; https://doi.org/10.3390/catal14120846 - 23 Nov 2024
Viewed by 450
Abstract
The application of advance oxidation processes (AOPs) based on the activation of peroxymonosulfate (PMS) is a great concern for wastewater treatment. Herein, ZrO2-3DG was constructed using a hydrothermal method for the degradation of tetracycline (TC) with PMS. The defective ZrO2 [...] Read more.
The application of advance oxidation processes (AOPs) based on the activation of peroxymonosulfate (PMS) is a great concern for wastewater treatment. Herein, ZrO2-3DG was constructed using a hydrothermal method for the degradation of tetracycline (TC) with PMS. The defective ZrO2-3DG materials were also prepared with plasma treatment. SEM and XPS results show that the ZrO2-3DG composite and the corresponding defective materials were successfully fabricated. The ZrO2 particles are distributed uniformly on the substrate material. Plasma can induce defects on the composite materials and create highly active sites. TC degradation results show that the ZrO2-3DG/PMS system can achieve a degradation efficiency of 92.9% for TC. The influences of defects on materials, light irradiation and sulfate accumulation were investigated. It has been found that defects can induce an inhibiting effect on the degradation process, which can be tuned by plasma time. The defective ZrO2-3DG/PMS system exhibits excellent resistance to the accumulation of sulfate, even showing enhanced degradation performances in specific conditions. The light irradiation has led to a higher degradation efficiency with the accumulation of sulfate compared with a dark environment. These findings give great guidance to the application of the ZrO2-3DG/PMS system for environmental protection. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM and EDS-mapping images for (<b>a</b>–<b>f</b>) ZrO<sub>2</sub>-3DG and (<b>g</b>–<b>l</b>) defective D-ZrO<sub>2</sub>-3DG (60 s).</p>
Full article ">Figure 2
<p>(<b>a</b>) XPS survey and high-resolution XPS of; (<b>b</b>) C 1s, (<b>c</b>) O 1s and (<b>d</b>) Zr 3d for pristine ZrO<sub>2</sub>-3DG and D-ZrO<sub>2</sub>-3DG (60 s).</p>
Full article ">Figure 3
<p>Degradation of TC in dark environment with (<b>a</b>) PMS; (<b>b</b>) ZrO<sub>2</sub>-3DG/PMS system.</p>
Full article ">Figure 4
<p>Degradation of TC with defective D-ZrO<sub>2</sub>-3DG/PMS system in dark environment under plasma-treated conditions (<b>a</b>) 60 s, (<b>b</b>) 180 s and (<b>c</b>) 300 s.</p>
Full article ">Figure 5
<p>Degradation of TC with non-defective and D-ZrO<sub>2</sub>-3DG/PMS system under visible light irradiation with PMS addition of (<b>a</b>) 0.02 mmol, (<b>b</b>) 0.2 mmol, (<b>c</b>) 1.0 mmol and (<b>d</b>) 3.0 mmol.</p>
Full article ">Figure 6
<p>The changes in the apparent rate constant (kapp) of TC degradation with non-defective and D-ZrO<sub>2</sub>-3DG/PMS system under visible light irradiation, PMS of (<b>a</b>) 0.02 mmol; (<b>b</b>) 0.2 mmol; (<b>c</b>) 1.0 mmol; (<b>d</b>) 3.0 mmol.</p>
Full article ">Figure 7
<p>Degradation of TC and the corresponding changes in the apparent rate constant (kapp) with non-defective and D-ZrO<sub>2</sub>-3DG/PMS system in the additional SO<sub>4</sub><sup>2−</sup> accumulation of (<b>a</b>,<b>b</b>) 10 mM, (<b>c</b>,<b>d</b>) 30 mM and (<b>e</b>,<b>f</b>) 50 mM in a dark environment.</p>
Full article ">Figure 8
<p>Degradation of TC with non-defective and defective ZrO<sub>2</sub>-3DG/PMS system in the additional SO<sub>4</sub><sup>2−</sup> accumulation of (<b>a</b>) 10 mM, (<b>c</b>) 30 mM and (<b>e</b>) 50 mM under visible light irradiation. (<b>b</b>,<b>d</b>,<b>f</b>) The changes in the apparent rate constant (kapp) for the above processes.</p>
Full article ">
14 pages, 3347 KiB  
Article
Efficient Degradation of Tetracycline by Peroxymonosulfate Activated with Ni-Co Bimetallic Oxide Derived from Bimetallic Oxalate
by Qi Zhang, Mingling Yu, Hang Liu, Jin Tang, Xiaolong Yu, Haochuan Wu, Ling Jin and Jianteng Sun
Toxics 2024, 12(11), 816; https://doi.org/10.3390/toxics12110816 - 14 Nov 2024
Viewed by 531
Abstract
In this work, NiCo2O4 was synthesized from bimetallic oxalate and utilized as a heterogeneous catalyst to active peroxymonosulfate (PMS) for the degradation of tetracycline (TC). The degradation efficiency of TC (30 mg/L) in the NiCo2O4 + PMS [...] Read more.
In this work, NiCo2O4 was synthesized from bimetallic oxalate and utilized as a heterogeneous catalyst to active peroxymonosulfate (PMS) for the degradation of tetracycline (TC). The degradation efficiency of TC (30 mg/L) in the NiCo2O4 + PMS system reached 92.4%, with NiCo2O4 exhibiting satisfactory reusability, stability, and applicability. Radical trapping test and electron paramagnetic resonance (EPR) results indicated that SO4•−, •OH, O2•−, and 1O2 were the dominating reactive oxygen species (ROS) for TC degradation in the NiCo2O4 + PMS system. Seven intermediates were identified, and their degradation pathways were proposed. Toxicity assessment using T.E.S.T software (its version is 5.1.1.0) revealed that the identified intermediates had lower toxicity compared to intact TC. A rice seed germination test further confirmed that the NiCo2O4 + PMS system effectively degraded TC into low-toxicity or non-toxic products. In conclusion, NiCo2O4 shows promise as a safe and efficient catalyst in advanced oxidation processes (AOPs) for the degradation of organic pollutants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The XRD pattern (<b>a</b>), FTIR spectrum (<b>b</b>), SEM (<b>c</b>), and EDS (<b>d</b>) of the obtained NiCo<sub>2</sub>O<sub>4</sub>.</p>
Full article ">Figure 2
<p>The comparison of different catalysts for TC degradation (<b>a</b>) and degradation kinetics (<b>b</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 3
<p>The degradation efficiency of TC in the presence of different scavengers: TBA (<b>a</b>); EtOH (<b>b</b>); <span class="html-italic">p</span>-BQ (<b>c</b>); and FFA (<b>d</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 4
<p>EPR spectrum of the NiCo<sub>2</sub>O<sub>4</sub> + PMS system with the addition of DMPO or TEMP: (<b>a</b>) SO<sub>4</sub><sup>−</sup> and OH; (<b>b</b>) O<sub>2</sub><sup>−</sup>; (<b>c</b>) <sup>1</sup>O<sub>2</sub>.</p>
Full article ">Figure 5
<p>The reusability, stability, and applicability of NiCo<sub>2</sub>O<sub>4</sub> (<b>a</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM. The XRD of the fresh and used of NiCo<sub>2</sub>O<sub>4</sub> (<b>b</b>). The degradation efficiency of other organic contaminants in the NiCo<sub>2</sub>O<sub>4</sub> + PMS system (<b>c</b>).</p>
Full article ">Figure 6
<p>The effect of experimental factors on TC degradation: pH (<b>a</b>); PMS concentration (<b>b</b>); NiCo<sub>2</sub>O<sub>4</sub> dosage (<b>c</b>); and TC concentration (<b>d</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 7
<p>The germination of rice seedlings in before reaction solution and after reaction solution.</p>
Full article ">
12 pages, 2690 KiB  
Article
Perborate Activated Peroxymonosulfate Process for Improving the Coagulation Efficiency of Microcystis aeruginosa by Polymeric Aluminum Chloride
by Fan Chen, Lu Li, Shunfan Qiu, Shiyang Chen, Lingfang Yang, Lin Deng and Zhou Shi
Molecules 2024, 29(22), 5352; https://doi.org/10.3390/molecules29225352 - 14 Nov 2024
Viewed by 525
Abstract
In this study, the sodium perborate (SP)-activated peroxymonosulfate (PMS) process was used to enhance the coagulation efficiency of cyanobacteria with polymeric aluminum chloride (PAC), aiming to efficiently mitigate the impact of algal blooms on the safety of drinking water production. The optimal concentrations [...] Read more.
In this study, the sodium perborate (SP)-activated peroxymonosulfate (PMS) process was used to enhance the coagulation efficiency of cyanobacteria with polymeric aluminum chloride (PAC), aiming to efficiently mitigate the impact of algal blooms on the safety of drinking water production. The optimal concentrations of SP, PMS, and PAC were determined by evaluating the removal rate of OD680 and zeta potential of the algae. Experimental results demonstrated that the proposed ternary PMS/SP/PAC process achieved a remarkable OD680 removal efficiency of 95.2%, significantly surpassing those obtained from individual treatments with PMS (19.5%), SP (5.2%), and PAC (42.1%), as well as combined treatments with PMS/PAC (55.7%) and PMS/SP (28%). The synergistic effect of PMS/SP/PAC led to the enhanced aggregation of cyanobacteria cells due to a substantial reduction in their zeta potential. Flow cytometry was performed to investigate cell integrity before and after treatment with PMS/SP/PAC. Disinfection by-products (DBPs) (sodium hypochlorite disinfection) of the algae-laden water subsequent to PMS/SP/PAC treatment declined by 57.1%. Moreover, microcystin-LR was completely degraded by PMS/SP/PAC. Electron paramagnetic resonance (EPR) analysis evidenced the continuous production of SO4, •OH, 1O2, and O2, contributing to both cell destruction and organic matter degradation. This study highlighted the significant potential offered by the PMS/SP/PAC process for treating algae-laden water. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of different processes on cyanobacteria removal (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 2
<p>Effect of SP and PMS dosage on cyanobacteria removal using PMS/SP/PAC treatment (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, PAC dosage: 10 mg/L).</p>
Full article ">Figure 3
<p>Effect of PAC dosage on cyanobacteria removal using PMS/SP/PAC treatment (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM).</p>
Full article ">Figure 4
<p>Cell integrity of algae treated with (<b>a</b>) control, (<b>b</b>) PAC, (<b>c</b>) PMS/PAC, and (<b>d</b>) PMS/SP/PAC (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 5
<p>DBP yield from different treatments of cyanobacteria under chlorination in DI water (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 6
<p>Microcystin-LR elimination using different treatment processes in the removal of cyanobacteria (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 7
<p>The EPR spectra of (<b>a</b>) DMPO−OH and DMPO−SO<sub>4</sub> adducts, (<b>b</b>) DMPO−<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> <mrow> <mo>•</mo> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> adducts, and (<b>c</b>) TEMP−<sup>1</sup>O<sub>2</sub> adducts (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L, [DMPO] = [TEMP] = 100 mM).</p>
Full article ">
13 pages, 1506 KiB  
Article
Impact of Metal Ions, Peroxymonosulfate (PMS), and pH on Sulfolane Degradation by Pressurized Ozonation
by Nasim Zare and Gopal Achari
Water 2024, 16(22), 3162; https://doi.org/10.3390/w16223162 - 5 Nov 2024
Viewed by 573
Abstract
This study investigated the degradation of sulfolane using pressurized ozonation under varying initial concentrations and the influence of different catalysts and peroxymonosulfate activation methods on the degradation efficiency. Initial sulfolane concentrations of 1 mg L−1, 20 mg L−1, and [...] Read more.
This study investigated the degradation of sulfolane using pressurized ozonation under varying initial concentrations and the influence of different catalysts and peroxymonosulfate activation methods on the degradation efficiency. Initial sulfolane concentrations of 1 mg L−1, 20 mg L−1, and 100 mg L−1 were tested over 120 min, revealing a degradation efficiency of 73%, 41%, and 18%, respectively. The addition of various metal ions (Zn2+, Mg2+, Cu2+, Ni2+, and Co2+) demonstrated that only zinc and magnesium enhanced degradation, with zinc achieving a 92% removal efficiency and magnesium achieving 86%. Different doses of magnesium and zinc were further tested, showing optimal degradation at specific concentrations. The combination of PMS with ozonation was explored, revealing that zinc activation did not significantly enhance degradation, while NaOH activation achieved near-total degradation, with a 100 mg L−1 NaOH concentration. Varying PMS concentrations indicated that altering pH was more effective than changing PMS dosage. Finally, the impact of pH changes in both reverse osmosis water and tap water matrices confirmed that higher pH levels significantly improved degradation efficacy, achieving up to 98% removal with NaOH concentrations of 50 mg L−1 in reverse osmosis water. These results suggest that optimizing pH and catalyst type are critical for enhancing sulfolane degradation in pressurized ozonation systems. Full article
(This article belongs to the Section Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Sulfolane degradation with different initial concentrations (initial C<sub>sulfolane</sub> = 1, 20, and 100 mg L<sup>−1</sup>; ozone flow rate = 0.6 L min<sup>−1</sup>; system pressure = 20 psi; and ozone percentage = 100%).</p>
Full article ">Figure 2
<p>Impact of homogenous metals on sulfolane degradation (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, ozone flow rate = 0.6 L min<sup>−1</sup>, system pressure = 20 psi, and ozone percentage = 100%).</p>
Full article ">Figure 3
<p>Impact of different initial zinc concentrations on sulfolane degradation (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>; C<sub>zinc</sub> = 2, 10, and 50 mg L<sup>−1</sup>; ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">Figure 4
<p>Sulfolane degradation with PMS activated with zinc (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, C<sub>PMS</sub> = 70 mg L<sup>−1</sup>, C<sub>zinc</sub> = 10 mg L<sup>−1</sup>, ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">Figure 5
<p>Sulfolane degradation with PMS activated with NaOH (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, C<sub>PMS</sub> = 70 mg L<sup>−1</sup>, C<sub>NaOH</sub> = 10 and 100 mg L<sup>−1</sup>, ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">Figure 6
<p>Impact of initial PMS dosage on sulfolane degradation (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, C<sub>PMS</sub> = 10 and 70 mg L<sup>−1</sup>, C<sub>NaOH</sub> = 10 mg L<sup>−1</sup>, ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">Figure 7
<p>Impact of NaOH on sulfolane degradation in reverse osmosis (RO) water (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">Figure 8
<p>Impact of NaOH on sulfolane degradation with tap water (initial C<sub>sulfolane</sub> = 1 mg L<sup>−1</sup>, ozone concentration = 2.86 mg L<sup>−1</sup>).</p>
Full article ">
16 pages, 3173 KiB  
Article
Activation of Persulfate by NiFe-Layered Double Hydroxides Toward Efficient Degradation of Doxycycline in Water
by Jie Chen, Xiaojun Tang, Jing Wang, Shiming Bi, Yinhan Lin and Zhujian Huang
Catalysts 2024, 14(11), 782; https://doi.org/10.3390/catal14110782 - 4 Nov 2024
Viewed by 817
Abstract
In recent years, the excessive use and improper disposal of antibiotics have led to their pervasive presence in the environment, resulting in significant antibiotic pollution. To address this pressing issue, the present study synthesized nickel–iron-layered double hydroxides (NiFe-LDHs) with varying molar ratios using [...] Read more.
In recent years, the excessive use and improper disposal of antibiotics have led to their pervasive presence in the environment, resulting in significant antibiotic pollution. To address this pressing issue, the present study synthesized nickel–iron-layered double hydroxides (NiFe-LDHs) with varying molar ratios using a hydrothermal method, employing these LDHs as catalysts for the oxidative degradation of doxycycline, with peroxymonosulfate (PMS) serving as the oxidant. X-ray diffraction analysis confirmed that the synthesized NiFe-LDHs exhibited a hexagonal crystal structure characteristic of layered double hydroxides. Experimental results demonstrated that the catalytic efficiency of NiFe-LDHs increased with both the dosage of the catalyst and the concentration of PMS, achieving a high degradation efficiency for doxycycline at a catalyst concentration of 0.5 g/L. Furthermore, the catalytic performance was notably effective across a range of pH conditions, with the highest degradation efficiency being observed at a Ni–Fe molar ratio of 3:1. The activation of PMS by NiFe-LDHs for the catalytic degradation of pollutants primarily occurs through singlet oxygen (1O2), superoxide radicals (O2·), and sulfate radicals (SO4·). The study also proposed three potential degradation pathways for doxycycline, indicating that the final degradation products have lower environmental toxicity. This research offers novel approaches and methodologies for the treatment of antibiotic-contaminated wastewater. Full article
(This article belongs to the Section Environmental Catalysis)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of NiFe-LDHs; (<b>b</b>) TEM image of the 3: 1 NiFe-LDHs.</p>
Full article ">Figure 2
<p>High resolution of XPS (<b>a</b>) C 1s (<b>b</b>) O1s (<b>c</b>) Ni 2p and (<b>d</b>) Fe 2p for NiFe-LDHs.</p>
Full article ">Figure 3
<p>(<b>a</b>) Adsorption-desorption isotherms and (<b>b</b>) pore size distribution curves.</p>
Full article ">Figure 4
<p>Effect of PMS dosage on the degradation of doxycycline hydrochloride (dosed PMS quantity: 0 g/L; 0.02 g/L; 0.04 g/L; 0.08 g/L; 0.10 g/L).</p>
Full article ">Figure 5
<p>Catalyst dosing effects on the degradation of doxycycline hydrochloride in wastewater treatment (dosed catalyst quantity: 0.1 g/L; 0.3 g/L; 0.4 g/L; 0.5 g/L).</p>
Full article ">Figure 6
<p>Effect of pH value of solution on the degradation of doxycycline hydrochloride (pH = 3; pH = 4; pH = 6; pH = 7; pH = 8; pH = 10).</p>
Full article ">Figure 7
<p>Effect of catalysts on the degradation of doxycycline hydrochloride (1:1 NiFe-LDH; 2:1 NiFe-LDH; 3:1 NiFe-LDH; 4:1 NiFe-LDH).</p>
Full article ">Figure 8
<p>ESR spectra of the NiFe-LDHs activated PMS system under the condition of adding (<b>a</b>,<b>b</b>) DMPO (5,5-Dimethyl-1-pyrroline N-oxide) and (<b>c</b>) TEMP (2,2,6,6-Tetramethylpiperidine) as spin-trapping agents. (<b>d</b>) Removal efficiency of doxycycline in NiFe-LDHs/PMS with scavenger.</p>
Full article ">Figure 9
<p>Degradation pathway of doxycycline hydrochloride.</p>
Full article ">
16 pages, 6360 KiB  
Article
Magnetic MgFeO@BC Derived from Rice Husk as Peroxymonosulfate Activator for Sulfamethoxazole Degradation: Performance and Reaction Mechanism
by Tong Liu, Chen-Xuan Li, Xing Chen, Yihan Chen, Kangping Cui and Qiang Wei
Int. J. Mol. Sci. 2024, 25(21), 11768; https://doi.org/10.3390/ijms252111768 - 1 Nov 2024
Viewed by 647
Abstract
Heterogeneous Mg-Fe oxide/biochar (MgFeO@BC) nanocomposites were synthesized by a co-precipitation method and used as biochar-based catalysts to activate peroxymonosulfate (PMS) for sulfamethoxazole (SMX) removal. The optimal conditions for SMX degradation were examined as follows: pH 7.0, MgFeO@BC of 0.4 g/L, PMS concentration of [...] Read more.
Heterogeneous Mg-Fe oxide/biochar (MgFeO@BC) nanocomposites were synthesized by a co-precipitation method and used as biochar-based catalysts to activate peroxymonosulfate (PMS) for sulfamethoxazole (SMX) removal. The optimal conditions for SMX degradation were examined as follows: pH 7.0, MgFeO@BC of 0.4 g/L, PMS concentration of 0.6 mM and SMX concentration of 10.0 mg/L at 25 ℃. In the MgFeO@BC/PMS system, the removal efficiency of SMX was 99.0% in water within 40 min under optimal conditions. In the MgFeO@BC/PMS system, the removal efficiencies of tetracycline (TC), cephalexin (CEX), ciprofloxacin (CIP), 4-chloro-3-methyl phenol (CMP) and SMX within 40 min are 95.3%, 98.4%, 98.2%, 97.5% and 99.0%, respectively. The radical quenching experiments and electron spin resonance (ESR) analysis suggested that both non-radical pathway and radical pathway advanced SMX degradation. SMX was oxidized by sulfate radicals (SO4•−), hydroxyl radicals (•OH) and singlet oxygen (1O2), and SO4•− acted as the main active species. MgFeO@BC exhibits a higher current density, and therefore, a higher electron migration rate and redox capacity. Due to the large number of available binding sites on the surface of MgFeO@BC and the low amount of ion leaching during the catalytic reaction, the system has good anti-interference ability and stability. Finally, the intermediates of SMX were detected. Full article
(This article belongs to the Section Materials Science)
Show Figures

Figure 1

Figure 1
<p>SEM images of BC (<b>a</b>), Fe<sub>3</sub>O<sub>4</sub>@BC (<b>b</b>), MgO@BC (<b>c</b>) and MgFeO@BC (<b>d</b>).</p>
Full article ">Figure 2
<p>HRTEM images (<b>a</b>,<b>b</b>), SAED (<b>c</b>), EDX (<b>d</b>) and C, Fe, Mg, N, O element distribution (<b>e</b>–<b>g</b>).</p>
Full article ">Figure 3
<p>XRD patterns (<b>a</b>), FT-IR spectra (<b>b</b>), Raman spectra (<b>c</b>), room-temperature magnetization curve (<b>d</b>), nitrogen adsorption–desorption isotherms (<b>e</b>) and pore size distribution (<b>f</b>) of catalysts.</p>
Full article ">Figure 4
<p>The adsorption efficiencies of SMX in various reaction systems. Reaction conditions: [SMX]<sub>0</sub> = 10.0 mg/L, [catalyst]<sub>0</sub> = 0.4 g/L, initial pH = 7.0 and T = 25 ℃.</p>
Full article ">Figure 5
<p>SMX degradation efficiencies (<b>a</b>), and <span class="html-italic">k</span><sub>obs</sub> in different systems (<b>b</b>). Reaction conditions: [SMX]<sub>0</sub> = 10.0 mg/L, [catalyst]<sub>0</sub> = 0.4 g/L, [PMS]<sub>0</sub> = 0.6 mM, initial pH = 7.0 and T = 25 ℃.</p>
Full article ">Figure 6
<p>Effects of different parameters on SMX removal in MgFeO@BC/PMS system: (<b>a</b>) catalyst dosage, (<b>b</b>) PMS concentration, (<b>c</b>) solution pH, (<b>d</b>) SMX concentration. (Conditions: pH = 7.0; [SMX]<sub>0</sub> = 10.0 mg/L; [catalyst] = 0.4 g/L; [PMS] = 0.6 mM; T = 25 ℃.)</p>
Full article ">Figure 7
<p>Zeta potentials of MgFeO@BC catalyst.</p>
Full article ">Figure 8
<p>SMX degradation efficiency under different scavengers (<b>a</b>–<b>d</b>). (Conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [MgFeO@BC]<sub>0</sub> = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; T = 25 ℃.)</p>
Full article ">Figure 9
<p>ESR signals of (<b>a</b>) TEMP-<sup>1</sup>O<sub>2</sub>, and (<b>b</b>) DMPO-X. (Conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [MgFeO@BC]<sub>0</sub> = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; T = 25 ℃; [TEMP] = [DMPO] = 10.0 mM.)</p>
Full article ">Figure 10
<p>XPS spectra of full-range survey (<b>a</b>), C 1s (<b>b</b>), Fe 2p (<b>c</b>) for MgFeO@BC; CV curves of catalysts (<b>d</b>).</p>
Full article ">Figure 11
<p>Possible pathways for degradation of SMX in MgFeO@BC/PMS system.</p>
Full article ">Figure 12
<p>Proposed mechanism of SMX degradation in MgFeO@BC/PMS system.</p>
Full article ">Figure 13
<p>Effects of Cl<sup>−</sup> (<b>a</b>), HCO<sub>3</sub><sup>−</sup> (<b>b</b>), NO<sub>3</sub><sup>−</sup> (<b>c</b>) and HA (<b>d</b>) on SMX degradation. (Conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [MgFeO@BC]<sub>0</sub> = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; T = 25 ℃.)</p>
Full article ">Figure 14
<p>The degradation of SMX in different aqueous substrates (<b>a</b>), the removal of various pollutants in MgFeO@BC/PMS system (<b>b</b>), reusability of MgFeO@BC after 5 consecutive cycles (<b>c</b>), and the concentration of leached ions (<b>d</b>). (Conditions: [substrate]<sub>0</sub> = 10.0 mg/L; [MgFeO@BC]<sub>0</sub> = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; T = 25 ℃.)</p>
Full article ">
13 pages, 11209 KiB  
Article
Natural Vanadium–Titanium Magnetite Activated Peroxydisulfate and Peroxymonosulfate for Acid Orange II Degradation: Different Activation Mechanisms and Influencing Factors
by Zheng Zhang, Libin Zhao, Jingyuan Tian, Shaojie Ren and Wei Zhang
Water 2024, 16(21), 3109; https://doi.org/10.3390/w16213109 - 30 Oct 2024
Viewed by 522
Abstract
Persulfate-based advanced oxidation processes have emerged as a promising approach for the degradation of organic pollutants in aqueous environments due to their ability to generate sulfate radicals (SO4−·) within catalytic systems. In this study, peroxydisulfate (PDS) and peroxymonosulfate (PMS) were [...] Read more.
Persulfate-based advanced oxidation processes have emerged as a promising approach for the degradation of organic pollutants in aqueous environments due to their ability to generate sulfate radicals (SO4−·) within catalytic systems. In this study, peroxydisulfate (PDS) and peroxymonosulfate (PMS) were investigated with the natural vanadium–titanium magnetite (VTM) as the activator for the degradation of acid orange II. The degradation efficiency increased with higher dosages of VTM or persulfate (both PDS and PMS) at lower concentrations (below 10 mM). However, excessive PMS (higher than 10 mM) in the PMS/VTM system led to the self-consumption of free radicals, significantly inhibiting the degradation of acid orange II. The VTM-activated PDS or PMS maintained an effective degradation of acid orange II in a wide pH range (3~11), suggesting remarkable pH stability. The SO4−· was the main active species in the PDS/VTM system, while hydroxyl radical (·OH) also contributed significantly to the PMS/VTM system. In addition, PMS exhibited better thermal stability during VTM activation. Coexisting ions in an aqueous environment such as bicarbonate (HCO3), carbonate (CO32–), and hydrogen phosphate (HPO42–) had obvious effects on persulfate activation. Our study systematically investigated the different activation processes and influencing factors associated with PDS and PMS when the natural VTM was used as a catalyst, thereby providing new insights into the persulfate-mediated degradation of organic pollutants in aqueous environments. Full article
(This article belongs to the Topic Advanced Oxidation Processes for Wastewater Purification)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM and EDS mapping images of the VTM. (<b>a</b>–<b>d</b>) Surface morphology of VTM at different resolutions. EDS results of different elements on the VTM surface: (<b>e</b>) merged image, (<b>f</b>) Fe, (<b>g</b>) V, (<b>h</b>) Ti, (<b>i</b>) C, and (<b>j</b>) O elements.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns <b>of</b> the VTM. (<b>b</b>) Nitrogen adsorption−desorption isotherms; inset image was the pore size distribution of the VTM.</p>
Full article ">Figure 3
<p>XPS spectra of the VTM before and after acid orange II degradation tests: (<b>a</b>) full-range scan, (<b>b</b>) Fe 2p spectra, (<b>c</b>) O 1s spectra, and (<b>d</b>) C 1s spectra.</p>
Full article ">Figure 4
<p>Effect of VTM dosage on the acid orange II degradation in (<b>a</b>) VTM/PDS system and (<b>b</b>) VTM/PMS system. Effect of (<b>c</b>) PDS and (<b>d</b>) PMS concentrations on the acid orange II degradation efficiency.</p>
Full article ">Figure 5
<p>Effect of different initial pH on the degradation efficiency of acid orange II in the (<b>a</b>) VTM/PDS system and (<b>b</b>) VTM/PMS system. The inset image is the final pH after a 60 min reaction. Effect of different temperatures on the degradation efficiency of acid orange II in the (<b>c</b>) VTM/PDS system and (<b>d</b>) VTM/PMS system. The inset image is the color change in the acid orange II after a 60 min reaction.</p>
Full article ">Figure 6
<p>(<b>a</b>) Effect of EtOH on acid orange II degradation in the VTM/PDS reaction system. (<b>b</b>) Effect of TBA on acid orange II degradation in the VTM/PDS reaction system. (<b>c</b>) The influence of EtOH on acid orange II degradation in the VTM/PMS reaction system. (<b>d</b>) The influence of TBA on acid orange II degradation in the VTM/PMS reaction system.</p>
Full article ">Figure 7
<p>UV-vis analysis of acid orange II in the VTM/PDS reaction system (<b>a</b>) and VTM/PMS reaction system (<b>b</b>).</p>
Full article ">
14 pages, 10730 KiB  
Article
WS2-Assisted Electrochemical Activation of Peroxymonosulfate for Eliminating Organic Pollutant in Water
by Wenxuan Du, Xiren Xia, Zhen Li, Fuzhen Liu and Yin Xu
Catalysts 2024, 14(11), 763; https://doi.org/10.3390/catal14110763 - 29 Oct 2024
Viewed by 689
Abstract
Advanced oxidation process based on heterogeneous activation of peroxymonosulfate (PMS) has received significant attention in wastewater remediation. Herein, a facile and effective electrochemical method was introduced in a tungsten sulfide (WS2)-activated PMS process for the removal of a typical azo dye [...] Read more.
Advanced oxidation process based on heterogeneous activation of peroxymonosulfate (PMS) has received significant attention in wastewater remediation. Herein, a facile and effective electrochemical method was introduced in a tungsten sulfide (WS2)-activated PMS process for the removal of a typical azo dye Acid Orange 7 (AO7) in aqueous solution. It was found that the electrochemical activation could remarkably promote the removal of organic pollutants by coupling with WS2/PMS system. The elimination of AO7 in the electro-assisted WS2-activated PMS (E/WS2/PMS) system achieved 95.8% of AO7 removal in 30 min, with the optimal conditions of 1.0 g/L WS2, 1.0 mM PMS, current density of 1.0 mA/cm2 and initial pH of 6.5. Based on quenching experiments and EPR techniques, mechanistic studies confirmed that hydroxyl radical (OH) and singlet oxygen (1O2) are the primary reactive oxygen species for the oxidation of pollutants. In addition, the influences of pH, WS2 dosage, PMS concentration, current density, common anions and humic acid on the AO7 removal are also investigated in detail. Furthermore, the system exhibited resistance to aqueous matrices, verifying the accepted applicability in real water (i.e., Yangtze River water and Shahu Lake water). In summary, this study demonstrates a green system for the effective removal of contaminants in water, holding significant implications for practical application. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern of WS<sub>2</sub> sample.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) SEM images of WS<sub>2</sub> and (<b>d</b>) EDS of WS<sub>2</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Removal of AO7 in different systems and (<b>b</b>) the reaction rate constant of AO7 removal of different systems. Experimental conditions: [AO7]<sub>0</sub> = 10 mg/L, [WS<sub>2</sub>]<sub>0</sub> = 1.0 g/L, [PMS]<sub>0</sub> = 1.0 mM, current density = 1.0 mA/cm<sup>2</sup> and pH<sub>0</sub> = 6.5.</p>
Full article ">Figure 4
<p>Influence of (<b>a</b>) WS<sub>2</sub> dosage, (<b>b</b>) PMS concentration, (<b>c</b>) current density and (<b>d</b>) initial pH on the removal of AO7. Experimental conditions: [PMS]<sub>0</sub> = 1.0 mM, [WS<sub>2</sub>]<sub>0</sub> = 1.0 g/L, current intensity = 1.0 mA/cm<sup>2</sup>, pH<sub>0</sub> = 6.5.</p>
Full article ">Figure 5
<p>(<b>a</b>) Influence of different anions of 5 mM, (<b>b</b>) HA and (<b>c</b>) real water on the removal of AO7. Experimental conditions: [AO7]<sub>0</sub> = 10 mg/L, [WS<sub>2</sub>]<sub>0</sub> = 1.0 g/L, [PMS]<sub>0</sub> = 1.0 mM, current density = 1.0 mA/cm<sup>2</sup> and pH<sub>0</sub> = 6.5.</p>
Full article ">Figure 6
<p>EPR spectra for the detection of (<b>a</b>) <sup>•</sup>OH in the presence of DMPO. (<b>b</b>) <sup>1</sup>O<sub>2</sub> in the presence of TEMP. (<b>c</b>) <math display="inline"><semantics> <msup> <msubsup> <mi mathvariant="normal">O</mi> <mn>2</mn> <mo>•</mo> </msubsup> <mo>−</mo> </msup> </semantics></math> in the presence of DMPO (in ethanol solution) in different systems.</p>
Full article ">Figure 7
<p>(<b>a</b>) The inhibition of different quenchers on the removal of AO7 and (<b>b</b>) their reaction rate constants of AO7 removal. Experimental conditions: [AO7]<sub>0</sub> = 10 mg/L, [WS<sub>2</sub>]<sub>0</sub> = 1.0 g/L, [PMS]<sub>0</sub> = 1.0 mM, current density = 1.0 mA/cm<sup>2</sup>, [TBA/MeOH] = 600 mM, [L-His] = 10 mM and pH<sub>0</sub> = 6.5.</p>
Full article ">Figure 8
<p>Schematic diagram for the mechanism of the E/WS<sub>2</sub>/PMS system.</p>
Full article ">Figure 9
<p>(<b>a</b>) Stability test of WS<sub>2</sub> and (<b>b</b>) the adsorption of AO7 by used WS<sub>2</sub>. Experimental conditions: [AO7]<sub>0</sub> = 10 mg/L, [WS<sub>2</sub>]<sub>0</sub> = 1.0 g/L, [PMS]<sub>0</sub> = 1.0 mM and initial pH = 6.5.</p>
Full article ">Figure 10
<p>(<b>a</b>) XRD patterns and (<b>b</b>) full range of XPS spectrum, (<b>c</b>) W 4f; (<b>d</b>) S 2p spectra of fresh and used WS<sub>2</sub>.</p>
Full article ">Figure 11
<p>A possible degradation pathway of AO7 in the E/WS<sub>2</sub>/PMS system.</p>
Full article ">Figure 12
<p>Developmental toxicity of AO7 and the degradation products.</p>
Full article ">
14 pages, 6899 KiB  
Article
Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups
by Tong Liu, Chen-Xuan Li, Xing Chen, Yihan Chen, Kangping Cui and Qiang Wei
Int. J. Mol. Sci. 2024, 25(21), 11582; https://doi.org/10.3390/ijms252111582 - 29 Oct 2024
Cited by 2 | Viewed by 756
Abstract
In this work, rice-husk-derived biochar (RBC) was synthesized by using simple one-step pyrolysis strategies and served as catalysts to activate peroxymonosulfate (PMS) for degrading sulfamethoxazole (SMX). When the annealing temperature (T) = 800 °C, RBC800 exhibits the typical hardwood structure with several [...] Read more.
In this work, rice-husk-derived biochar (RBC) was synthesized by using simple one-step pyrolysis strategies and served as catalysts to activate peroxymonosulfate (PMS) for degrading sulfamethoxazole (SMX). When the annealing temperature (T) = 800 °C, RBC800 exhibits the typical hardwood structure with several micropores and mesoporous. Furthermore, RBC800 obtains more defect sites than RBC600, RBC700, and RBC900. In the RBC800/PMS system, the removal rate of the SMX reached 92.0% under optimal conditions. The kinetic reaction rate constant (kobs) of the RBC800/PMS system was 0.009 min−1, which was about 1.50, 1.28, and 4.50 times that of the RBC600/PMS (kobs = 0.006 min−1), RBC700/PMS (kobs = 0.007 min−1), and RBC900/PMS (kobs = 0.002 min−1) systems, respectively. In the RBC800/PMS system, sulfate radical (SO4•−) is the main active species. Compared with other active sites, the hydroxyl group (C-OH) on the surface of RBC800 interacts more strongly with PMS, which is more likely to promote the stretching of the O-O bond of the PMS, thus breaking into the activated state and significantly reducing the activation energy required for reaction. The degradation intermediates of SMX were speculated, and the toxicity analysis was conducted. Generally, this work reveals in depth the interaction between reactive sites of biochar-based catalysts and PMS at the molecular level. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>d</b>) SEM images of RBC<sub>T</sub>; (<b>e</b>) EDS elemental content, and (<b>f</b>) element mappings of RBC<sub>800</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>) HRTEM, and (<b>b</b>) SAED pattern of RBC<sub>800</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD patterns, (<b>b</b>) FTIR spectra, (<b>c</b>) Nitrogen adsorption–desorption isotherms and pore structure (the inset), and (<b>d</b>) Raman spectra of RBC<sub>T</sub>.</p>
Full article ">Figure 4
<p>The adsorption rate (<b>a</b>) and degradation rate (<b>b</b>) of SMX in various reaction systems. Reaction conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [RBC<sub>T</sub>] = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C.</p>
Full article ">Figure 5
<p>Effects of ROSs scavengers on the SMX degradation in the RBC<sub>800</sub>/PMS system. Reaction conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [RBC<sub>800</sub>] = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; MeOH = 0.5 M; TBA = 0.5 M; FFA = 0.5 M; <span class="html-italic">p</span>-BQ = 20.0 mM; Reaction temperature = 25 °C.</p>
Full article ">Figure 6
<p>ESR signals of (<b>a</b>) TEMP-<sup>1</sup>O<sub>2</sub> and (<b>b</b>) DMPO-<sup>•</sup>OH and DMPO-SO<sub>4</sub><sup>•−</sup>. (Conditions: [SMX]<sub>0</sub> = 10.0 mg/L; [RBC<sub>800</sub>]<sub>0</sub> = 0.4 g/L; [PMS]<sub>0</sub> = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C; [TEMP] = [DMPO] = 10.0 mM).</p>
Full article ">Figure 7
<p>XPS spectra of full-range survey (<b>a</b>), and O 1s (<b>b</b>) of RBC<sub>800</sub>.</p>
Full article ">Figure 8
<p>The optimization structures of PMS adsorption on different sites and the corresponding transition state. (<b>a</b>) C/PMS, (<b>b</b>) COOH/PMS, (<b>c</b>) C-OH/PMS and (<b>d</b>) C=O/PMS.</p>
Full article ">Figure 9
<p>Proposed mechanism of SMX degradation in the RBC<sub>800</sub>/PMS system.</p>
Full article ">Figure 10
<p>Bioaccumulation factor (<b>a</b>), and developmental toxicity (<b>b</b>) of SMX and its degradation byproducts.</p>
Full article ">
14 pages, 3850 KiB  
Article
Cobalt-Doped Carbon Nitride for Efficient Removal of Microcystis aeruginosa via Peroxymonosulfate Activation
by Wen Yan, Chuqiao Li, Yunjuan Meng, Yao Yue, Teer Wen, Jiafeng Ding and Hangjun Zhang
Toxins 2024, 16(11), 455; https://doi.org/10.3390/toxins16110455 - 24 Oct 2024
Viewed by 716
Abstract
Heterogeneous persulfate activation is an advanced technology for treating harmful algae in drinking water sources, while it remains a significant hurdle in the efficient management of cyanobacterial blooms. In this study, super-dispersed cobalt-doped carbon nitride (2CoCN) was prepared to activate peroxymonosulfate (PMS) for [...] Read more.
Heterogeneous persulfate activation is an advanced technology for treating harmful algae in drinking water sources, while it remains a significant hurdle in the efficient management of cyanobacterial blooms. In this study, super-dispersed cobalt-doped carbon nitride (2CoCN) was prepared to activate peroxymonosulfate (PMS) for simultaneous Microcystis aeruginosa inhibition and microcystin (MC-LR) degradation. When the initial PMS and 2CoCN concentrations were 0.3 g/L and 0.4 g/L, respectively, the efficiency of algal cell removal reached 97% in 15 min, and the degradation of MC-LR reached 96%. Analyses by SEM, TEM, and EEM spectra revealed that the reaction led to changes in algal cell morphology, damage to the cell membrane and cell wall, and the diffusion of thylakoid membranes and liposomes. The activities of antioxidant enzymes (superoxide dismutase and catalase) and antioxidants (glutathione) in algal cells generally increased, and the content of malondialdehyde increased, indicating severe damage to the cell membrane. Radical capture experiments confirmed that singlet oxygen (1O₂) was the key species destroying algal cells in the 2CoCN/PMS system. The 2CoCN/PMS system was effective in removing M. aeruginosa within a wide pH range (3–9), and 2CoCN had good reusability. Additionally, three degradation products of MC-LR were identified by LC–MS/MS analysis, and a possible mechanism for the inactivation of M. aeruginosa and the degradation of MC-LR was proposed. In conclusion, this study pioneered the 2CoCN/PMS system for inhibiting M. aeruginosa and degrading microcystin, aiming to advance water purification and algae removal technology. Full article
Show Figures

Figure 1

Figure 1
<p>Variations of algal removal efficiency (<b>a</b>) and MC-LR concentrations (<b>b</b>) in the 2CoCN/PMS system. (C/C₀ represents the concentration of MC-LR during the reaction divided by its initial concentration.) Error bars represent the standard deviations (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2
<p>The SEM images of surface morphology of <span class="html-italic">M. aeruginosa</span>: before (<b>a</b>) and after reaction (<b>b</b>); the TEM images: before (<b>c</b>) and after reaction (<b>d</b>); chlorophyll a spontaneous fluorescence intensity: before (<b>e</b>) and after reaction (<b>f</b>).</p>
Full article ">Figure 3
<p>The excitation emission matrix spectrum of EOM and IOM; EOM: 0 min (<b>a</b>), 15 min (<b>b</b>), 30 min (<b>c</b>); IOM: 0 min (<b>d</b>), 15 min (<b>e</b>), 30 min (<b>f</b>).</p>
Full article ">Figure 4
<p>Response surface (<b>a</b>) and 3D contour plot (<b>b</b>) of interactive effect between the different factors for the removal of algal cells; typical time courses for algae removal in different pH (<b>c</b>); recyclability experiment (<b>d</b>).</p>
Full article ">Figure 5
<p>Oxidative stress indicators after 30 min of reaction: SOD (<b>a</b>), CAT (<b>b</b>), GSH (<b>c</b>), and MDA (<b>d</b>); (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Radical scavenging experiment of algae removal in 2CoCN/PMS system (<b>a</b>); © reaction contains approximately 0.60 mg·L<sup>−1</sup>of chlorophyll <span class="html-italic">a</span>, 0.15 g·L<sup>−1</sup>of 2CoCN and 0.30 mM of PMS and 20 mM of the scavengers; EPR spectrum: <sup>1</sup>O<sub>2</sub> (<b>b</b>), O<sub>2</sub><sup>•−</sup> (<b>c</b>), SO<sub>4</sub><sup>•−</sup> and <sup>•</sup>OH (<b>d</b>).</p>
Full article ">Figure 7
<p>The possible degradation pathway of MC-LR in 2CoCN/PMS system.</p>
Full article ">Figure 8
<p>Possible mechanisms of homogeneous/heterogeneous ROS−flocculation process for <span class="html-italic">M. aeruginosa</span> removal in 2CoCN/PMS system.</p>
Full article ">
14 pages, 4950 KiB  
Article
Construction of Co-Modified MXene/PES Catalytic Membrane for Effective Separation and Degradation of Tetracycline Antibiotics in Aqueous Solutions
by Xiaojie Cheng, Xiaojun Qin, Runxue Zhao, Jiamin Chen, Xia Zheng, Ke Liu and Meixuan Xin
Molecules 2024, 29(21), 4995; https://doi.org/10.3390/molecules29214995 - 22 Oct 2024
Viewed by 626
Abstract
The application of antibiotics has advanced modern medicine significantly. However, the abuse and discharge of antibiotics have led to substantial antibiotic residues in water, posing great harm to natural organisms and humans. To address the problem of antibiotic degradation, this study developed a [...] Read more.
The application of antibiotics has advanced modern medicine significantly. However, the abuse and discharge of antibiotics have led to substantial antibiotic residues in water, posing great harm to natural organisms and humans. To address the problem of antibiotic degradation, this study developed a novel catalytic membrane by depositing Co catalysts onto MXene nanosheets and fabricating the polyethersulfone composite (Co@MXene/PES) using vacuum-assisted self-assembly. The dual role of MXene as both a carrier for Co atoms and an enhancer of interlayer spacing led to improved flux and catalytic degradation capabilities of the membrane. Experimental results confirmed that the Co@MXene/PES membrane effectively degraded antibiotics through peroxymonosulfate activation, achieving up to 95.51% degradation at a cobalt concentration of 0.01 mg/mL. The membrane demonstrated excellent antibacterial properties, minimal flux loss after repeated use, and robust anti-fouling performance, making it a promising solution for efficient antibiotic removal and stable water treatment. Full article
(This article belongs to the Special Issue Nano Environmental Materials II)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) FTIR spectra of MXene and Co-MXene; (<b>b</b>) XRD patterns of MXene and Co-MXene.</p>
Full article ">Figure 2
<p>SEM images of pure MXene (<b>a1</b>) and the optimal ratio of Co-MXene (<b>a2</b>); TEM images of pure MXene (<b>b1</b>) and the optimal ratio of Co-MXene (<b>b2</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) Cross-sectional SEM images of pure MXene membrane and the optimal ratio Co@MXene/PES membrane; (<b>b</b>) EDS mapping of the optimal ratio Co@MXene/PES membrane.</p>
Full article ">Figure 4
<p>(<b>a</b>) 2D AFM diagram of pure MXene membrane and the optimal ratio Co@MXene/PES membrane; (<b>b</b>) 3D AFM plot of pure MXene membrane and the optimal ratio Co@MXene/PES membrane.</p>
Full article ">Figure 5
<p>(<b>a</b>) The water contact angle of M0–M3; (<b>b</b>) Permeability of M0–M3; (<b>c</b>) M0–M3 pure retention for TC.</p>
Full article ">Figure 6
<p>(<b>a</b>) Removal rate of TC by different membranes; (<b>b</b>) Effect of PMS concentration on TC removal rate (pH = 6, T = 55 °C); (<b>c</b>) Effect of pH value on TC removal rate (PMS concentration = 30 mg, T = 55 °C); (<b>d</b>) Effect of temperature on TC removal rate (PMS concentration = 30 mg, pH = 6).</p>
Full article ">Figure 7
<p>(<b>a</b>) Colonial growth of <span class="html-italic">S. aureus</span> on membrane and (<b>b</b>) schematic diagram of antibacterial mechanism.</p>
Full article ">Figure 8
<p>Stability and cyclic tests of membranes: (<b>a</b>) Digital photos of two-dimensional membranes after being immersed in solutions with different pH values for 1 day and 7 days (M0, M1, M2 and M3, from left to right); (<b>b</b>) The permeability and TC removal changes of the membrane under four cycles.</p>
Full article ">Figure 9
<p>EPR spectra of (<b>a</b>) <sup>•</sup>OH, SO<sub>4</sub><sup>•−</sup> and (<b>b</b>) O<sub>2</sub><sup>•−</sup>.</p>
Full article ">Figure 10
<p>The degradation and separation mechanism of Co@MXene/PES membrane.</p>
Full article ">Figure 11
<p>The fabrication steps of Co@MXene/PES membrane.</p>
Full article ">
18 pages, 12639 KiB  
Article
Iron–Cobalt Bimetallic Metal–Organic Framework-Derived Carbon Materials Activate PMS to Degrade Tetracycline Hydrochloride in Water
by Qin Liu, Huali Zhang, Kanghui Zhang, Jinxiu Li, Jiaheng Cui and Tongshan Shi
Water 2024, 16(20), 2997; https://doi.org/10.3390/w16202997 - 21 Oct 2024
Viewed by 792
Abstract
Organic pollutants entering water bodies lead to severe water pollution, posing a threat to human health. The activation of persulfate advanced oxidation processes using carbon materials derived from MOFs as substrates can efficiently treat wastewater contaminated with organic pollutants. This research uses NH [...] Read more.
Organic pollutants entering water bodies lead to severe water pollution, posing a threat to human health. The activation of persulfate advanced oxidation processes using carbon materials derived from MOFs as substrates can efficiently treat wastewater contaminated with organic pollutants. This research uses NH2-MIL-101(Fe) as a substrate, doped with Fe2+ and Co2+, to prepare Fe/Co-CNs through a one-step carbonization method. The surface morphology, pore structure, and chemical composition of Fe/Co-CNs were investigated using characterization techniques such as XRD, SEM, TEM, XPS, FT-IR, BET, and Raman. A comparative study was conducted on the performance of catalysts with different Fe/Co ratios in activating PMS for the degradation of organic pollutants, as well as the effects of various influencing factors (the dosage of Fe/Co-CNs, the amount of peroxymonosulfate (PMS), the initial pH of the solution, the TC concentration, and inorganic anions) on the catalyst’s activation of persulfate for TC degradation. Through radical quenching experiments and post-degradation XPS analysis, the active radicals in the FeCo-CNs/PMS system were investigated to explain the possible mechanism of TC degradation in the Fe/Co-CNs/PMS system. The results indicate that Fe/Co-CNs-2 (with a Co2+ doping amount of 20%) achieves a degradation rate of 93.34% for TC (tetracycline hydrochloride) within 30 min when activating PMS, outperforming other Co2+ doping amounts. In addition, singlet oxygen (1O2) is the main reactive species in the reaction system. Full article
(This article belongs to the Section Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Preparation of Fe/Co-CNs.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of Fe/Co-CNs-2; (<b>b</b>) XRD patterns of Fe/Co-CNs-1, Fe/Co-CNs-2, Fe/Co-CNs-3, and Fe/Co-CNs-4.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>f</b>) SEM images of Fe/Co-CNs-2; (<b>g</b>–<b>k</b>) Elemental mapping of Fe/Co-CNs-2.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>f</b>) TEM images of Fe/Co-CNs-2.</p>
Full article ">Figure 5
<p>XPS spectra of Fe/Co-CNs-2: (<b>a</b>) Survey, (<b>b</b>) C 1s, (<b>c</b>) Fe 2p, and (<b>d</b>) Co 2p.</p>
Full article ">Figure 6
<p>FTIR patterns of Fe/Co-CNs-2.</p>
Full article ">Figure 7
<p>N<sub>2</sub> adsorption/desorption isotherm of Fe/Co-CNs-2.</p>
Full article ">Figure 8
<p>Raman spectra of Fe/Co-CNs-2 and Fe/Co-CNs-4.</p>
Full article ">Figure 9
<p>(<b>a</b>) Degradation efficiency of TC by Fe-CNs-7 and Fe/Co-CN-activated PMS (<b>b</b>) Degradation efficiency of TC by Fe/Co-CNs-2-activated PMS/PDS.</p>
Full article ">Figure 10
<p>Effect of catalyst dosage on TC degradation by the Fe/Co-CNs-2/PMS system.</p>
Full article ">Figure 11
<p>Effect of PMS consumption on TC degradation by the Fe/Co-CNs-2/PMS system.</p>
Full article ">Figure 12
<p>Effects of initial concentration on the degradation of TC.</p>
Full article ">Figure 13
<p>Effect of initial pH of solution on degradation of TC.</p>
Full article ">Figure 14
<p>Effects of different anions on the degradation of TC in FeCo-CNs-2/PMS system: (<b>a</b>) Cl<sup>−</sup>, (<b>b</b>) CO<sub>3</sub><sup>2−</sup>, (<b>c</b>) HCO<sup>3−</sup>, (<b>d</b>) HPO<sub>4</sub><sup>2−</sup>, (<b>e</b>) SO<sub>4</sub><sup>2−</sup>.</p>
Full article ">Figure 15
<p>Recycling efficiency of TC degradation by the Fe/Co-CNs-2/PMS system.</p>
Full article ">Figure 16
<p>(<b>a</b>) Effects of 10 mM of different radical scavengers and (<b>b</b>) 100 mM of different radical scavengers on the degradation of TC in the Fe/Co-CNs-2/PMS system.</p>
Full article ">Figure 17
<p>XPS spectrum of Fe/Co-CNs-2 after degradation: (<b>a</b>) Fe 2p and (<b>b</b>) Co 2p.</p>
Full article ">
Back to TopTop