[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,717)

Search Parameters:
Keywords = remediation efficiency

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 1969 KiB  
Article
Bioleaching of Metal-Polluted Mine Tailings: A Comparative Approach Between Ex Situ Slurry-Phase Stirred Reactors Versus In Situ Electrokinetic Percolation
by Irene Acosta Hernández, Martín Muñoz Morales, Francisco Jesús Fernández Morales, Luis Rodríguez Romero and José Villaseñor Camacho
Appl. Sci. 2024, 14(24), 11756; https://doi.org/10.3390/app142411756 - 17 Dec 2024
Viewed by 222
Abstract
This work compares two technologies for the remediation of metal-polluted mine tailings based on lab-scale bioleaching experiments performed in (a) conventional agitated slurry-phase reactors and (b) in situ electrokinetic percolation. While ex situ bioleaching in agitated reactors has been widely studied, only a [...] Read more.
This work compares two technologies for the remediation of metal-polluted mine tailings based on lab-scale bioleaching experiments performed in (a) conventional agitated slurry-phase reactors and (b) in situ electrokinetic percolation. While ex situ bioleaching in agitated reactors has been widely studied, only a few previous works have studied the in situ option that couples bioleaching and electrokinetics. Real mine tailings from an abandoned sphalerite mine in southern Spain were used. The leaching medium was externally generated in a bioreactor using an autochthonous acidophilic culture and then added to tailings in batch experiments. This medium enabled metal leaching from mine tailings without the stringent operating conditions required by a classic bioleaching process. Metal removal efficiencies and kinetic rate constants after 15 d of treatments were calculated. Additionally, advantages or disadvantages between the two methods were discussed. The results for the innovative EK-percolation method showed rates and efficiencies that were comparable to, and in some cases better than, those achieved with conventional stirred slurry systems. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic experimental setup and procedure.</p>
Full article ">Figure 2
<p>Efficiencies (%) in metal solubilization from mine tailings using a soil slurry agitated system, using the BL medium (continuous lines) and an iron-free TK medium (dashed lines, reference tests).</p>
Full article ">Figure 3
<p>Efficiencies (%) in metal solubilization from tailings using an EK-percolation system, using as an anolyte the BL medium (continuous lines) or an iron-free TK medium (dashed lines, reference tests).</p>
Full article ">Figure 4
<p>Metal removal yields (%) at the end of experiments (AS: agitated slurry; EKP: EK-percolation tests; RT: reference test; BL: biologically produced leaching medium).</p>
Full article ">Figure 5
<p>Kinetic modeling of Mn leaching data: (<b>Left Column</b>): agitated slurry system; (<b>Right Column</b>): EK-percolation. Reference test (white points) and using BL medium (black points). Two models are plotted: (<b>a</b>) Top figures: considering the leaching rate controlled by a solid product layer diffusion (shrinking core model); (<b>b</b>) Bottom figures: considering the leaching rate controlled by a chemical reaction at the particles’ surface (shrinking particle model).</p>
Full article ">Figure 6
<p>Kinetic constant values for the leaching of all metals using BL medium (AS: agitated slurry; EKP: EK-percolation tests). The two proposed models have been considered: (a) Equation (1) and (b) Equation (2).</p>
Full article ">
13 pages, 5496 KiB  
Article
Sustainable Removal of Phenol Dye-Containing Wastewater by Composite Incorporating ZnFe2O4/Nanocellulose Photocatalysts
by Zan Li, Kun Gao, Wenrui Jiang, Jiao Xu and Pavel Lushchyk
Sustainability 2024, 16(24), 11023; https://doi.org/10.3390/su162411023 - 16 Dec 2024
Viewed by 328
Abstract
The escalating issue of phenol-containing wastewater necessitates the development of efficient and sustainable treatment methods. In this context, we present a novel composite photocatalyst comprising ZnFe2O4 (ZFO) nanoparticles supported on nanocellulose (NC), aimed at addressing this environmental challenge. The synthesis [...] Read more.
The escalating issue of phenol-containing wastewater necessitates the development of efficient and sustainable treatment methods. In this context, we present a novel composite photocatalyst comprising ZnFe2O4 (ZFO) nanoparticles supported on nanocellulose (NC), aimed at addressing this environmental challenge. The synthesis involved a facile hydrothermal method followed by the impregnation of ZFO nanoparticles onto the NC matrix. The morphology and structure of ZFO, NC, and ZFO/NC were investigated by TEM, SEM-EDX, UV–vis, FT-IR, XRD, and XPS analyses. ZFO, as a weakly magnetic semiconductor catalytic material, was utilized in photocatalytic experiments under magnetic field conditions. By controlling the electron spin states through the magnetic field, electron–hole recombination was suppressed, resulting in improved photocatalytic performance. The results demonstrated that 43% and 76% degradation was achieved after 120 min of irradiation due to ZFO and 0.5ZFO/NC treatment. Furthermore, the composite 0.5ZFO/NC demonstrated the highest photocatalytic efficiency, showing promising recyclability by maintaining its activity after three cycles of use. This study underscores the potential of the ZFO/NC composite for sustainable wastewater treatment, offering a promising avenue for environmental remediation. Full article
(This article belongs to the Special Issue Advanced Materials and Processes for Wastewater Treatment)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of NC, ZFO, and hybrid nanocomposites.</p>
Full article ">Figure 2
<p>FT-IR spectra of NC, ZFO, and ZFO/NC nanocomposites.</p>
Full article ">Figure 3
<p>(<b>a</b>) UV–visible spectra and (<b>b</b>) Tauc plots for the band gap of NC, ZFO, and ZFO/NC nanocomposites.</p>
Full article ">Figure 4
<p>XPS survey spectra of ZFO and ZFO/NC: (<b>a</b>) Fe 2p, (<b>b</b>) Zn 2p, (<b>c</b>) O 1s, (<b>d</b>) C 1s.</p>
Full article ">Figure 5
<p>Characterization of synthesized NC and nanocomposites. (<b>a</b>,<b>b</b>) denote TEM images for NC and ZFO/NC, respectively. (<b>c</b>,<b>d</b>) denote SEM images for NC and ZFO/NC, respectively.</p>
Full article ">Figure 6
<p>Elemental compositions of (<b>a</b>) ZFO and NC and (<b>b</b>) hybrid nanocomposites; (<b>c</b>) EDS mapping results for ZFO.</p>
Full article ">Figure 7
<p>(<b>a</b>) Degradation curves of phenol by NC, ZFO, and xZFO/NC (x = 0.1, 0.3, 0.5, and 0.7) in the absence of a magnetic field; (<b>b</b>) degradation curves of phenol by NC, ZFO, and xZFO/NC under magnetic field conditions; (<b>c</b>) comparison of the degradation efficiency of xZFO/NC in the absence of a magnetic field and in the presence of a magnetic field; (<b>d</b>) the percentage increase in the photodegradation rate of xZFO/NC after the addition of a magnetic field.</p>
Full article ">Figure 8
<p>(<b>a</b>) UV–visible absorption spectra of phenol and phenol under UV–vis; (<b>b</b>) absorption spectra of the degradation of phenol by 0.5ZFO/NC under the condition of a magnetic field as a function of time.</p>
Full article ">Figure 9
<p>Cyclic experiments on photocatalytic degradation of phenol by 0.5ZFO/NC.</p>
Full article ">Figure 10
<p>0.5ZFO/NC in the absence of a magnetic field and in the presence of a magnetic field (MF = magnetic field; NMF = no magnetic field): (<b>a</b>) photocurrent response density; (<b>b</b>) electrochemical impedance spectroscopy (CPE = Constant Phase Angle Element).</p>
Full article ">Figure 11
<p>Schematic of the mechanism of visible-light photocatalytic phenol degradation by NC, ZFO, and xZFO/NC.</p>
Full article ">
21 pages, 6307 KiB  
Article
Visible Light-Driven Direct Z-Scheme Ho2SmSbO7/YbDyBiNbO7 Heterojunction Photocatalyst for Efficient Degradation of Fenitrothion
by Liang Hao and Jingfei Luan
Molecules 2024, 29(24), 5930; https://doi.org/10.3390/molecules29245930 - 16 Dec 2024
Viewed by 216
Abstract
A highly versatile Z-scheme heterostructure, Ho2SmSbO7/YbDyBiNbO7 (HYO), was synthesized using an ultrasonic-assisted solvent thermal method. The HYO heterojunction, composed of dual A2B2O7 compounds, exhibits superior separation of photogenerated carriers due to its efficient [...] Read more.
A highly versatile Z-scheme heterostructure, Ho2SmSbO7/YbDyBiNbO7 (HYO), was synthesized using an ultrasonic-assisted solvent thermal method. The HYO heterojunction, composed of dual A2B2O7 compounds, exhibits superior separation of photogenerated carriers due to its efficient Z-scheme mechanism. The synergistic properties of Ho2SmSbO7 and YbDyBiNbO7, particularly the excellent visible light absorption, enable HYO to achieve exceptional photocatalytic performance in the degradation of fenitrothion (FNT). Specifically, HYO demonstrated an outstanding removal efficiency of 99.83% for FNT and a mineralization rate of 98.77% for total organic carbon (TOC) during the degradation process. Comparative analyses revealed that HYO significantly outperformed other photocatalysts, including Ho2SmSbO7, YbDyBiNbO7, and N-doped TiO2, achieving removal rates that were 1.10, 1.20, and 2.97 times higher for FNT, respectively. For TOC mineralization, HYO exhibited even greater enhancements, with rates 1.13, 1.26, and 3.37 times higher than those of the aforementioned catalysts. Additionally, the stability and durability of HYO were systematically evaluated, confirming its potential applicability in practical scenarios. Trapping experiments and electron paramagnetic resonance analyses were conducted to identify the active species generated by HYO, specifically hydroxyl radicals (•OH), superoxide anions (•O2), and holes (h+). This facilitated a comprehensive understanding of the degradation mechanisms and pathways associated with FNT. In conclusion, this study represents a substantial contribution to the advancement of efficient Z-scheme heterostructure and offers critical insights for the development of sustainable remediation approaches aimed at mitigating FNT contamination. Full article
(This article belongs to the Special Issue Advances in Oxide-Based Materials for (Photo)Catalysis)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD, (<b>b</b>) FTIR, and (<b>c</b>) Raman plots of Ho<sub>2</sub>SmSbO<sub>7</sub>, YbDyBiNbO<sub>7</sub>, and HYO.</p>
Full article ">Figure 2
<p>(<b>a</b>) TEM, (<b>b</b>) SEM, (<b>c</b>) HRTEM, (<b>d</b>) EDS layered, (<b>e</b>) EDS, and (<b>f</b>) EDS elemental mapping images of HYO.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) TEM, (<b>b</b>) SEM, (<b>c</b>) HRTEM, (<b>d</b>) EDS layered, (<b>e</b>) EDS, and (<b>f</b>) EDS elemental mapping images of HYO.</p>
Full article ">Figure 3
<p>The XPS spectrum of synthesized Ho<sub>2</sub>SmSbO<sub>7</sub>, YbDyBiNbO<sub>7</sub>, and HYO: (<b>a</b>) survey spectrum, (<b>b</b>–<b>g</b>) high-resolution spectra of Bi 4f, Ho 4d, Sm 3d, Yb 4d, Dy 4d, Nb 4d, Sb 3d, and O 1S, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–vis DRS and (<b>b</b>) corresponding plots of (<span class="html-italic">αhν</span>)<sup>1/2</sup> and <span class="html-italic">hν</span> for Ho<sub>2</sub>SmSbO<sub>7</sub>, YbDyBiNbO<sub>7</sub>, and HYO; (<b>c</b>) PL spectra, (<b>d</b>) TRPL spectra, (<b>e</b>) PC curves, and (<b>f</b>) EIS plots of Ho<sub>2</sub>SmSbO<sub>7</sub>, YbDyBiNbO<sub>7</sub>, and HYO.</p>
Full article ">Figure 5
<p>(<b>a</b>) Photodegradation, (<b>b</b>) kinetic curves, and (<b>c</b>) removal efficiencies and kinetic constants for FNT; (<b>d</b>) mineralization, (<b>e</b>) kinetic curves, and (<b>f</b>) mineralization efficiencies and kinetic constants for TOC.</p>
Full article ">Figure 6
<p>(<b>a</b>) Five consecutive tests on HYO for the degradation of FNT and the mineralization of TOC under VE; (<b>b</b>) XRD and (<b>c</b>) XPS patterns of the fresh and the used HYO.</p>
Full article ">Figure 7
<p>Impact of different radical scavengers on (<b>a</b>) FNT saturation, (<b>b</b>) removal efficiency of FNT, and (<b>c</b>) EPR spectrum for DMPO·O<sub>2</sub><sup>−</sup> and DMPO·OH over HYO.</p>
Full article ">Figure 8
<p>UPS spectra of Ho<sub>2</sub>SmSbO<sub>7</sub> and YbDyBiNbO<sub>7</sub> (the intersections of the black dash lines indicated by the black arrows indicated the onset (Ei) and cutoff (Ecutoff) binding energy).</p>
Full article ">Figure 9
<p>Plausible photodegradation mechanism of FNT with HYO as photocatalyst under VE.</p>
Full article ">Figure 10
<p>Viable photodegradation pathway for FNT under VE with HYO heterojunction as catalyst.</p>
Full article ">
19 pages, 9163 KiB  
Article
Synthesis and Characterization of MgO-Fe₂O₃/γ-Al₂O₃ Nanocomposites: Enhanced Photocatalytic Efficiency and Selective Anticancer Properties
by ZabnAllah M. Alaizeri, Hisham A. Alhadlaq, Saad Aldawood and Maqusood Ahamed
Catalysts 2024, 14(12), 923; https://doi.org/10.3390/catal14120923 - 14 Dec 2024
Viewed by 383
Abstract
In the present work, we achieved the fabrication of MgO-Fe2O3/γ-Al2O3 NCs using a deposition–coprecipitation process. XRD, TEM, and SEM with EDX, XPS, FTIR, and PL spectroscopy were applied to examine the physicochemical properties of the samples. [...] Read more.
In the present work, we achieved the fabrication of MgO-Fe2O3/γ-Al2O3 NCs using a deposition–coprecipitation process. XRD, TEM, and SEM with EDX, XPS, FTIR, and PL spectroscopy were applied to examine the physicochemical properties of the samples. XRD analysis confirmed the successful incorporation of γ-Al2O3, MgO, and Fe2O3 phases. TEM and SEM images indicate that the nanocomposites exhibited an agglomerated morphology with spherical shapes and particle sizes in the range of 6–12 nm. EDX and XPS spectra revealed a composition of MgO-Fe2O3/γ-Al2O3 NCs. FTIR spectra identified characteristic vibrational bands corresponding to the chemical bonds present in the samples, confirming their successful synthesis. PL analysis showed the reduced recombination rate of electron–hole pairs and enhanced charge separation efficiency, which are important factors for improved photocatalytic activity. Photocatalysis results show that the MgO-Fe2O3/γ-Al2O3 NCs exhibited significantly higher photocatalysis efficiencies of 87.5% for Rh B and 90.4% for MB after 140 min, compared to γ-Al2O3 NPs and Fe2O3/γ-Al2O3 NPs. In addition, prepared MgO-Fe2O3/γ-Al2O3 NCs demonstrated superior stability after six runs. Biochemical data showed that the MgO-Fe2O3/γ-Al2O3 NCs exhibited significant toxicity toward A549 cancer cells while displaying low toxicity toward IMR90 normal cells. The IC50 values (µg/mL ± SD) for γ-Al2O3 NPs, Fe2O3/γ-Al2O3 NPs, and MgO-Fe2O3/γ-Al2O3 NCs were 16.54 ± 0.8 µg/mL, 14.75 ± 0.4 µg/mL, and 11.40 ± 0.6 µg/mL, respectively. These results suggest that the addition of Fe2O3 and MgO to γ-Al2O3 not only enhances photocatalytic activity but also improves biocompatibility and anticancer properties. This study highlights that the MgO-Fe2O3/γ-Al2O3 NCs warrant further exploration of their potential applications in environmental remediation and biomedicine. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, and (<b>c</b>) MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 2
<p>TEM images, HRTEM images, and SAED analysis: (<b>a</b>–<b>c</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, (<b>d</b>–<b>f</b>) Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, and (<b>j</b>–<b>i</b>) MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 3
<p>SEM images: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, (<b>c</b>) MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs, and (<b>d</b>) EDX analysis of MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 4
<p>Elemental mapping of the distribution of MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs: (<b>a</b>) electron, (<b>b</b>) aluminum (Al), (<b>c</b>) iron (Fe), (<b>d</b>) magnesium (Mg), and (<b>e</b>) oxygen (O).</p>
Full article ">Figure 5
<p>XPS spectra: (<b>a</b>) XPS survey spectra and high-resolution XPS spectra of (<b>b</b>) Al 1p, (<b>c</b>) Fe 2p, (<b>d</b>) O 1 s, and (<b>e</b>) Mg 2p for <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, and MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 6
<p>FTIR spectra of the synthesized samples: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, and (<b>c</b>) MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 7
<p>Photoluminescence spectra of <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, and MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 8
<p>(<b>a</b>) UV-Vis absorption of the Rh dye solution, (<b>b</b>) plot (C<sub>t</sub>/C<sub>0</sub>) vs. irradiation time (min), (<b>c</b>) kinetics of the photocatalysis of Rh B solutions for the prepared samples, and (<b>d</b>) photocatalysis efficiency (D%) of the Rh B solution using the synthesized catalyst.</p>
Full article ">Figure 9
<p>(<b>a</b>) UV-Vis absorption of MB dye solution, (<b>b</b>) plot (C<sub>t</sub>/C<sub>0</sub>) vs. irradiation time (min), (<b>c</b>) kinetics of the photocatalysis of MB solutions for prepared samples, and (<b>d</b>) photocatalysis efficiency (D%) of MB solution using synthesized catalyst.</p>
Full article ">Figure 10
<p>The number of recycled Rh B and MB dye photocatalysis agents using the prepared MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs under UV irradiation for 140 min.</p>
Full article ">Figure 11
<p>Schematic diagram of the photoreaction mechanism of organic dyes using MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs.</p>
Full article ">Figure 12
<p>The percentage of viable cells after exposure to the <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NPs, or MgO-Fe<sub>2</sub>O<sub>3</sub>/<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> </semantics></math>-Al<sub>2</sub>O<sub>3</sub> NCs after 24 h: (<b>a</b>) A549 cells and (<b>b</b>) normal IMR90 cells. The symbol (*) indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05) between the treated sample and the control.</p>
Full article ">
15 pages, 1818 KiB  
Article
Application of Phosphate-Based Binders for the Stabilization and Solidification of Metal-Contaminated Soil: Mechanisms and Efficacy Evaluation
by Shiliang Xu, Ayesha Imtiyaz Cheema, Yunhui Zhang and Bin Dong
Toxics 2024, 12(12), 907; https://doi.org/10.3390/toxics12120907 - 13 Dec 2024
Viewed by 311
Abstract
At present, contamination due to toxic metals is a global concern. The management of problems caused by heavy metals relies on stabilization/solidification, which is the most effective technique for the control of metal pollution in soil. This study examined the immobilization efficiency of [...] Read more.
At present, contamination due to toxic metals is a global concern. The management of problems caused by heavy metals relies on stabilization/solidification, which is the most effective technique for the control of metal pollution in soil. This study examined the immobilization efficiency of various phosphate-based binders (Na3PO4, Na2HPO4, NaH2PO4), in addition to ordinary Portland cement (OPC), MgO, and CaO, for the stabilization of multi-metal-contaminated soils. Moreover, this study focused on the leachability of copper, nickel, zinc, lead, cadmium, and manganese (Cu, Ni, Zn, Pb, Cd, Mn, respectively) over different time periods and with different concentrations. Batch leaching experiments were conducted to determine the leaching ratios and percentages of the various metal concentrations, along with measuring the pH values of the leachates. Our results indicate that the use of OPC was validated due to its superior immobilization performance across all metals present in the soil, but particularly with regard to metals in high concentrations. This was due to the formation of stable hydroxides and the high pH values, which assisted in abating the metals’ solubility. Additionally, phosphate-based binders, despite being environmentally favorable, were found to be less effective, particularly for Pb and Cu, and the leaching results exceeded non-hazardous waste limits. MgO showed reasonable immobilization results but was less effective compared to OPC; on the other hand, CaO exhibited increased leaching over time. Therefore, the present research serves primarily to highlight that OPC is more suitable for soil remediation at industrial sites and in the construction of infrastructure. Meanwhile, phosphate-based binders are shown to be more appropriate for eco-friendly, non-load-bearing applications. Full article
(This article belongs to the Section Toxicity Reduction and Environmental Remediation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Unconfined compressive strength values of contaminated soils cured for 7 days and 28 days.</p>
Full article ">Figure 2
<p>Leachate pH values of contaminated soils treated for 7 days and 28 days.</p>
Full article ">Figure 3
<p>Leaching concentrations of metals in contaminated soils after 7 and 28 days. (<b>a</b>) Leached Pb concentration for various phosphate based binders (<b>b</b>) Leached Cu concentration for various phosphate based binders (<b>c</b>) Leached Ni concentration for various phosphate based binders (<b>d</b>) Leached Zn concentration for various phosphate based binders (<b>e</b>) Leached Cd concentration for various phosphate based binders (<b>f</b>) Leached Mn concentration for various phosphate based binders.</p>
Full article ">Figure 4
<p>Post-experimental characterization of contaminated soils after 28 days. (<b>a</b>) explains the Xrd results of the contaminated soil after 28 days, (<b>b</b>) explains the FTIR results of the contaminated soil after 28 days, (<b>c</b>) explains the TGA/DTG results of the contaminated soil after 28 days.</p>
Full article ">
16 pages, 2714 KiB  
Article
Treatment of Swine Wastewater Using the Domestic Microalga Halochlorella rubescens KNUA214 for Bioenergy Production and Carotenoid Extraction
by Yu-Hee Seo, Jeong-Mi Do, Ho-Seong Suh, Su-Bin Park and Ho-Sung Yoon
Appl. Sci. 2024, 14(24), 11650; https://doi.org/10.3390/app142411650 - 13 Dec 2024
Viewed by 395
Abstract
The management of swine wastewater (SW) presents significant environmental challenges, requiring solutions that combine effective treatment with resource recovery. This study highlights the dual role of microalgae in wastewater remediation and bioenergy production. H. rubescens KNUA214 was cultivated in media containing varying concentrations [...] Read more.
The management of swine wastewater (SW) presents significant environmental challenges, requiring solutions that combine effective treatment with resource recovery. This study highlights the dual role of microalgae in wastewater remediation and bioenergy production. H. rubescens KNUA214 was cultivated in media containing varying concentrations of diluted swine wastewater (DSW; 0%, 25%, 50%, and 100%). Cultivating with Blue Green-11 (BG-11) medium + 50% DSW maximized biomass growth, the chlorophyll content, and carotenoid production. Nutrient removal efficiency in 100% DSW over 8 days demonstrated reductions of 59.3% in total nitrogen, 67.7% in ammonia nitrogen, and 40.7% in total phosphorus, confirming the species’ capacity for effective wastewater treatment. The carotenoid analysis using HPLC revealed that astaxanthin, lutein, canthaxanthin, and beta-carotene exhibited the highest levels in BG-11 + 50% DSW. Furthermore, the biomass analyses confirmed its potential for bioenergy applications, with high calorific values and significant polyunsaturated fatty acid concentrations, enhancing its utility for bioenergy and biolubricant production. These findings position H. rubescens KNUA214 as an effective resource for integrating SW management with the sustainable production of high-value biochemicals, offering environmental and economic benefits. Full article
(This article belongs to the Special Issue Bioprocessing and Fermentation Technology for Biomass Conversion)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Light microscope image of <span class="html-italic">H. rubescens</span> KNUA214. (<b>b</b>) Phylogenetic relationship of <span class="html-italic">H. rubescens</span> KNUA214 and its closely related species based on 18S rRNA sequence data. Numbers at nodes indicate percentage values derived from 500-bootstrap-analysis samples. The scale bar represents differences in nucleotide sequences.</p>
Full article ">Figure 2
<p>(<b>a</b>) Optical density, (<b>b</b>) dry weight, and (<b>c</b>) chlorophyll <span class="html-italic">a</span> and (<b>d</b>) chlorophyll <span class="html-italic">b</span> concentrations of <span class="html-italic">H. rubescens</span> KNUA214 under different concentrations of DSW for 8 days. Microscopy images of <span class="html-italic">H. rubescens</span> KNUA214 in (<b>e</b>) BG-11 and (<b>f</b>) 100% DSW over an 8-day period.</p>
Full article ">Figure 3
<p>(<b>a</b>) Nutrient concentration and (<b>b</b>) percentage of removal efficiency in 100% DSW by <span class="html-italic">H. rubescens</span> KNUA214 over an 8-day period. Statistical significance between groups is denoted as follows: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>(<b>a</b>) Astaxanthin, (<b>b</b>) lutein, (<b>c</b>) zeaxanthin, (<b>d</b>) canthaxanthin, and (<b>e</b>) beta-carotene contents of <span class="html-italic">H. rubescens</span> KNUA214 cultivated under different concentrations of DSW on 4 and 8 days. Statistical significance between groups is denoted as follows: * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
12 pages, 4684 KiB  
Article
Efficient Photocatalytic Removal of Aqueous Ammonia Nitrogen by g-C3N4/CoP Heterojunctions Under Visible Light Illumination
by Dongxu Wang, Wanfeng Mao, Lihong Zhao, Duo Meng, Jiaqi Tang and Tengfei Wu
Nanomaterials 2024, 14(24), 1996; https://doi.org/10.3390/nano14241996 - 13 Dec 2024
Viewed by 440
Abstract
With the development of industry, agriculture, and aquaculture, excessive ammonia nitrogen mainly involving ionic ammonia (NH4+) and molecular ammonia (NH3) has inevitable access to the aquatic environment, posing a severe threat to water safety. Photocatalytic technology shows great [...] Read more.
With the development of industry, agriculture, and aquaculture, excessive ammonia nitrogen mainly involving ionic ammonia (NH4+) and molecular ammonia (NH3) has inevitable access to the aquatic environment, posing a severe threat to water safety. Photocatalytic technology shows great advantages for ammonia nitrogen removal, such as its efficiency, reusability, low cost, and environmental friendliness. In this study, CP (g-C3N4/CoP) composite materials, which exhibited high-efficiency ammonia nitrogen removal, were synthesized through a simple self-assembly method. For the optimal CP-10 (10% CoP) samples, the removal rate of ammonia nitrogen reached up to 94.8% within 80 min under visible light illumination. In addition, the nitrogen selectivity S(N2) is about 60% for all oxidative products. The high performance of the CP-10 photocatalysts can be ascribed to the effective separation and transmission of electron–hole pairs caused by their heterogeneous structure. This research has significance for the application of photocatalysis for the remediation of ammonia nitrogen wastewater. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of synthesized photocatalysts.</p>
Full article ">Figure 2
<p>(<b>a</b>) TEM image; (<b>b</b>–<b>d</b>) HRTEM images of CoP/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Co 2p, (<b>b</b>) P 2p, (<b>c</b>) C 1s, and (<b>d</b>) N 1s XPS spectra of CoP/g-C<sub>3</sub>N<sub>4</sub> photocatalysts.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV-vis DRS, (<b>b</b>) band gap energy, (<b>c</b>) photocurrent transient responses, and (<b>d</b>) EIS Nyquist plots.</p>
Full article ">Figure 5
<p>(<b>a</b>) Ammonia nitrogen removal rate of prepared photocatalysts and (<b>b</b>) apparent rate constants (k) for ammonia nitrogen removal rate with the prepared photocatalysts.</p>
Full article ">Figure 6
<p>The influences of (<b>a</b>) pH value and (<b>b</b>) ionic strength on the removal of ammonia nitrogen.</p>
Full article ">Figure 7
<p>(<b>a</b>) Ammonia nitrogen removal under different experimental conditions and (<b>b</b>) percentage of influence of different experimental conditions for ammonia nitrogen removal.</p>
Full article ">Figure 8
<p>(<b>a</b>) The removal of ammonia nitrogen over CP-10 photocatalysts in the presence of various scavengers, (<b>b</b>) Mott–Schottky plot of CP-10, (<b>c</b>) schematic band gap structures of prepared photocatalysts, (<b>d</b>) TEMPO-h<sup>+</sup> and (<b>e</b>) DMPO-·OH for S-CN samples, and (<b>f</b>) conversion of inorganic nitrogen in the process of ammonia nitrogen removal.</p>
Full article ">Figure 9
<p>Schematic diagram of the possible photocatalytic mechanism of ammonia nitrogen removal.</p>
Full article ">
19 pages, 4917 KiB  
Article
Life Cycle Assessment of Crude Oil-Contaminated Soil Treated by Low-Temperature Thermal Desorption and Its Beneficial Reuse for Soil Amendment
by Young Ho Song, Geon Yong Kim, Da Yeon Kim and Yong Woo Hwang
Sustainability 2024, 16(24), 10900; https://doi.org/10.3390/su162410900 (registering DOI) - 12 Dec 2024
Viewed by 402
Abstract
The effectiveness of thermal treatment technologies for the remediation of soils contaminated with heavy hydrocarbons has been extensively documented in the scientific literature. In general, high-concentration crude-oil-contaminated soil is treated with high-temperature thermal desorption (HTTD) in order to achieve high remediation efficiency. However, [...] Read more.
The effectiveness of thermal treatment technologies for the remediation of soils contaminated with heavy hydrocarbons has been extensively documented in the scientific literature. In general, high-concentration crude-oil-contaminated soil is treated with high-temperature thermal desorption (HTTD) in order to achieve high remediation efficiency. However, this process has the unintended consequence of destroying soil fertility. Low-temperature thermal desorption (LTTD) represents an alternative approach that has been developed with the objective of remediating heavily crude-oil-contaminated soil in a more rapid and cost-effective manner while simultaneously enhancing soil fertility. The thermal desorption unit (TDU) was employed using both LTTD and HTTD, operating at 300 °C and 500 °C, respectively, with a 30 min residence time in the kiln. The concentration of total petroleum hydrocarbons (TPH) in both the LTTD- and HTTD-treated soils was found to be less than 1% by weight, thereby below regulatory standards. The environmental impacts of both processes were assessed using the OpenLCA software version 2.0. The HTTD process exhibited a total abiotic depletion potential (ADP) impact of 1.63 × 10−4 MJ and a global warming potential (GWP) of 414 kg CO2-eq. In contrast, LTTD demonstrated lower impacts, with an ADP of 1.29 × 10−4 MJ and a GWP of 278 kg CO2-eq. The transition from HTTD to LTTD resulted in a notable reduction in ADP by 20.5% and in GWP by 32.9%. The application of LTTD-treated soil coated with coke or carbonized residues has been demonstrated to serve as an effective soil amendment, with the capacity to sequester approximately 50% of organic hydrocarbon contaminants. The results of this study illustrate the potential of LTTD for not only economical and rapid soil remediation but also the enhancement of soil quality through beneficial reuse. Full article
(This article belongs to the Section Soil Conservation and Sustainability)
Show Figures

Figure 1

Figure 1
<p>GC/FID chromatogram (C10–C40).</p>
Full article ">Figure 2
<p>TGA of crude-oil-contaminated soil.</p>
Full article ">Figure 3
<p>Thermal desorption.</p>
Full article ">Figure 4
<p>System boundary of contaminated soil thermal desorption treatment.</p>
Full article ">Figure 5
<p>LCA results for abiotic depletion potential (ADP) and global warming potential (GWP) in HTTD.</p>
Full article ">Figure 6
<p>LCA results for abiotic depletion potential (ADP) and global warming potential (GWP) in LTTD.</p>
Full article ">Figure 7
<p>GWP and ADP contribution reduction effect for each material/energy.</p>
Full article ">Figure 8
<p>Gas chromatography (GC) chromatograms of crude-oil-contaminated soil samples be-fore (contaminated soil) and after being treated at 300 °C and 500 °C for 30 min.</p>
Full article ">Figure 9
<p>Results of toxicity test by cultivating microbes.</p>
Full article ">Figure 10
<p>Seed germination test.</p>
Full article ">
16 pages, 1614 KiB  
Article
Biogenic ZnO Nanoparticles Effectively Alleviate Cadmium-Induced Stress in Durum Wheat (Triticum durum Desf.) Plants
by Eleonora Coppa, Giulia Quagliata, Samuela Palombieri, Chiara Iavarone, Francesco Sestili, Daniele Del Buono and Stefania Astolfi
Environments 2024, 11(12), 285; https://doi.org/10.3390/environments11120285 - 12 Dec 2024
Viewed by 296
Abstract
This study investigated the potential of biogenic ZnO nanoparticles (ZnO-NPs) to alleviate cadmium (Cd) toxicity in durum wheat plants exposed for 14 days to 25 μM CdSO4. By applying ZnO-NPs at two different concentrations (25 and 50 mg L−1), we [...] Read more.
This study investigated the potential of biogenic ZnO nanoparticles (ZnO-NPs) to alleviate cadmium (Cd) toxicity in durum wheat plants exposed for 14 days to 25 μM CdSO4. By applying ZnO-NPs at two different concentrations (25 and 50 mg L−1), we observed increased chlorophyll content, beneficially impacting the photosynthetic efficiency, and enhanced sulfur, zinc, and iron accumulation. Moreover, the ZnO-NP treatment reduced the Cd accumulation in shoots, mitigating leaf chlorosis and oxidative damage. This response was clearly mediated by the increased thiol and phytochelatin production, as well as the enhanced sulfate uptake rate, with TdSultr1.3 as the most responsive gene coding for high-affinity transporter to Cd stress. In conclusion, the application of biogenic ZnO-NPs appears to be a promising approach for reducing the uptake of heavy metals by plants. In addition, it could be successfully used in combination with contamination prevention measures and/or remediation of contaminated sites to remove and mitigate the harmful effects of Cd on the environment and human health. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of foliar application of ZnO-NPs at the concentration of 0, 25, and 50 ppm on shoot and root biomass (FW, fresh weight), shoot-to-root ratio (insert), and chlorophyll content, measured as SPAD units, of durum wheat plants grown in the absence (C) or presence (Cd) of 25 μM Cd. Data are reported as the mean of three biological replicates ± SD (n = 3). Different lower case letters indicate statistically significant differences among the growth conditions (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Oxidative damage expressed as changes in MDA accumulation in shoot (<b>A</b>) and root (<b>B</b>) tissues of durum wheat plants grown in the absence (C) or presence (Cd) of 25 μM Cd and treated foliarly with ZnO-NPs at the concentration of 0, 25, and 50 ppm. Statistics as in <a href="#environments-11-00285-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Effect of foliar application of ZnO-NPs at the concentration of 0, 25, and 50 ppm on non-protein thiols concentration in the shoot (<b>A</b>) and root (<b>B</b>) tissues of durum wheat plants grown in the absence (C) or presence (Cd) of 25 μM Cd. Statistics as in <a href="#environments-11-00285-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 4
<p>Changes in Cd, Zn, and Fe concentration (from left to right) in the shoot and root tissues of durum wheat plants grown in the absence (C) or presence (Cd) of 25 μM Cd and treated foliarly with ZnO-NPs at the concentration of 0, 25, and 50 ppm. The translocation rate, a measure of a plant’s ability to move Cd from roots to shoots, was calculated as the percentage ratio of shoot-to-root Cd concentration. Statistics as in <a href="#environments-11-00285-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>Effect of foliar application of ZnO-NPs at the concentration of 0, 25 and 50 ppm on total S concentration (<b>on the left</b>) in shoot (<b>A</b>) and root (<b>B</b>) tissues and on the relative expression levels by qRT-PCR of the genes encoding high-affinity sulfate transporters (<span class="html-italic">TdSultr1.1</span> and <span class="html-italic">TdSultr1.3</span>) (<b>on the right</b>) in roots of durum wheat plants grown in the absence (C) or presence (Cd) of 25 μM Cd. Statistics as in <a href="#environments-11-00285-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 6
<p>Correlation networks of durum wheat (<b>A</b>) shoots and (<b>B</b>) roots. The blue and red lines represent the positive and negative correlations, respectively. Abbreviations are as follows: Cd, cadmium concentration; Fe, iron concentration; Zn, zinc concentration; S, total S concentration; Thiols, non-protein thiols concentration; MDA, malondialdehyde concentration; TR, Cd translocation rate; and ZnO-NPs, the concentration of biogenic ZnO-NPs applied as foliar spray.</p>
Full article ">
36 pages, 14352 KiB  
Article
Novel Nanocomposites and Biopolymer-Based Nanocomposites for Hexavalent Chromium Removal from Aqueous Media
by Adina-Elena Segneanu, Ionela Amalia Bradu, Mihaela Simona Calinescu (Bocanici), Gabriela Vlase, Titus Vlase, Daniel-Dumitru Herea, Gabriela Buema, Maria Mihailescu and Ioan Grozescu
Polymers 2024, 16(24), 3469; https://doi.org/10.3390/polym16243469 - 12 Dec 2024
Viewed by 487
Abstract
Designing new engineered materials derived from waste is essential for effective environmental remediation and reducing anthropogenic pollution in our economy. This study introduces an innovative method for remediating metal-contaminated water, using two distinct waste types: one biowaste (eggshell) and one industrial waste (fly [...] Read more.
Designing new engineered materials derived from waste is essential for effective environmental remediation and reducing anthropogenic pollution in our economy. This study introduces an innovative method for remediating metal-contaminated water, using two distinct waste types: one biowaste (eggshell) and one industrial waste (fly ash). We synthesized three novel, cost-effective nanoadsorbent types, including two new tertiary composites and two biopolymer-based composites (specifically k-carrageenan and chitosan), which targeted chromium removal from aqueous solutions. SEM analysis reveals that in the first composite, EMZ, zeolite, and magnetite nanoparticles are successfully integrated into the porous structure of the eggshell. In the second composite (FMZ), fly ash and magnetite particles are similarly loaded within the zeolite pores. Each biopolymer-based composite is derived by incorporating the corresponding tertiary composite (FMZ or EMZ) into the biopolymer framework. Structural modifications of the eggshell, zeolite, chitosan, and k-carrageenan resulted in notable increases in specific surface area, as confirmed by BET analysis. These enhancements significantly improve chromium adsorption efficiency for each adsorbent type developed. The adsorption performances achieved are as follows: EMZ (89.76%), FMZ (84.83%), EMZCa (96.64%), FMZCa (94.87%), EMZC (99.64%), and FMZC (97.67%). The findings indicate that chromium adsorption across all adsorbent types occurs via a multimolecular layer mechanism, which is characterized as spontaneous and endothermic. Desorption studies further demonstrate the high reusability of these nanomaterials. Overall, this research underscores the potential of utilizing waste materials for new performant engineered low-cost composites and biocomposites for environmental bioremediation applications. Full article
(This article belongs to the Special Issue Advances in Sustainable Polymeric Materials, 3rd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of composites, k-carrageenan-based nanocomposites, and chitosan-based nanocomposite preparation: EMZ/EMZC/EMZCa (<b>a</b>), and FMZ/FMZC/FMZCa (<b>b</b>).</p>
Full article ">Figure 2
<p>The nitrogen adsorption–desorption isotherms (<b>a</b>); pore distribution (<b>b</b>) for all adsorbents.</p>
Full article ">Figure 3
<p>XRD spectra of all adsorbent types prepared and their components (<b>a</b>), chitosan, carrageenan, FMZ, FMZCa, FMZC (<b>b</b>), and EMZ, EMZCa and EMZC (<b>c</b>).</p>
Full article ">Figure 4
<p>SEM images of FZM (<b>a</b>), FMZCa (<b>b</b>), FMZC (<b>c</b>), EMZ (<b>d</b>), EMZCa (<b>e</b>), and EMZC (<b>f</b>).</p>
Full article ">Figure 5
<p>FT-IR spectra of all adsorbent types prepared and their components (<b>a</b>), chitosan, carrageenan, FMZ, FMZCa, and FMZC (<b>b</b>), EMZ, EMZCa and EMZC (<b>c</b>).</p>
Full article ">Figure 6
<p>VSM of FMZ:FMZCa:FMZC (<b>a</b>) and EMZ:EMZCa:EMZC (<b>b</b>).</p>
Full article ">Figure 7
<p>TGA thermograms of EMZ (<b>a</b>), FMZ (<b>b</b>), EMZCa (<b>c</b>), FMZCa (<b>d</b>), EMZC (<b>e</b>), and FMZC (<b>f</b>).</p>
Full article ">Figure 8
<p>Chromium removal efficiency (<b>a</b>) and adsorption capacity (<b>b</b>) as a function of adsorbent mass.</p>
Full article ">Figure 9
<p>Relationship between chromium (<b>a</b>) initial concentration and chromium removal efficiency (%) and (<b>b</b>) initial concentration and adsorption capacity (mg/g).</p>
Full article ">Figure 10
<p>Relationship between chromium (<b>a</b>) pH and chromium removal efficiency (%) and (<b>b</b>) pH and adsorption capacity (mg/g).</p>
Full article ">Figure 11
<p>Relationship between chromium (<b>a</b>) time and chromium removal efficiency (%) and (<b>b</b>) time and adsorption capacity (mg/g).</p>
Full article ">Figure 12
<p>Relationship between chromium (<b>a</b>) temperature and chromium removal efficiency (%) and (<b>b</b>) temperature and adsorption capacity (mg/g).</p>
Full article ">Figure 13
<p>Removal efficiency and contact time relationship for all four adsorbents.</p>
Full article ">Figure 14
<p>FT-IR spectrum of FMZ, EMZ, FMZCa, EMZCa, FMZC, and EMZC after adsorption.</p>
Full article ">Figure 15
<p>SEM images of FZM (<b>a</b>), EMZ (<b>b</b>), FMZCa (<b>c</b>), EMZCa (<b>d</b>), FMZC (<b>e</b>), and EMZC (<b>f</b>) after adsorption.</p>
Full article ">Figure 16
<p>EDX spectrum of EMZ (<b>a</b>), FMZ (<b>b</b>), EMZCa (<b>c</b>), FMZCa (<b>d</b>), EMZC (<b>e</b>), and FMZC (<b>f</b>) after adsorption.</p>
Full article ">Figure 17
<p>Elemental composition (%) determined via EDX analysis for EMZCa (<b>a</b>); FMZCa (<b>b</b>); EMZC (<b>c</b>); FMZC (<b>d</b>); EMZ (<b>e</b>), and FMZ (<b>f</b>).</p>
Full article ">Figure 18
<p>Comparative DTG curves of all prepared adsorbents before/after adsorption: EMZCa (<b>a</b>); FMZCa (<b>b</b>), EMZC (<b>c</b>), FMZC (<b>d</b>), EMZ (<b>e</b>), and FMZ (<b>f</b>).</p>
Full article ">Figure 19
<p>Schematic representation of the adsorption mechanism of prepared adsorbents: FMZ/FMCa (<b>a</b>), FMZ/FMZC (<b>b</b>) EMZ/EMZCa (<b>c</b>), and EMZ/EMZC (<b>d</b>).</p>
Full article ">Figure 19 Cont.
<p>Schematic representation of the adsorption mechanism of prepared adsorbents: FMZ/FMCa (<b>a</b>), FMZ/FMZC (<b>b</b>) EMZ/EMZCa (<b>c</b>), and EMZ/EMZC (<b>d</b>).</p>
Full article ">Figure 20
<p>(<b>a</b>) Relationship between desorption rate and time. (<b>b</b>) Reusability of all prepared adsorbents.</p>
Full article ">
30 pages, 6284 KiB  
Article
A Biorefinery Approach Integrating Lipid and EPS Augmentation Along with Cr (III) Mitigation by Chlorella minutissima
by Sonia Choudhary, Mansi Tiwari and Krishna Mohan Poluri
Cells 2024, 13(24), 2047; https://doi.org/10.3390/cells13242047 - 11 Dec 2024
Viewed by 333
Abstract
The quest for cleaner and sustainable energy sources is crucial, considering the current scenario of a steep rise in energy consumption and the fuel crisis, exacerbated by diminishing fossil fuel reserves and rising pollutants. In particular, the bioaccumulation of hazardous substances like trivalent [...] Read more.
The quest for cleaner and sustainable energy sources is crucial, considering the current scenario of a steep rise in energy consumption and the fuel crisis, exacerbated by diminishing fossil fuel reserves and rising pollutants. In particular, the bioaccumulation of hazardous substances like trivalent chromium has not only disrupted the fragile equilibrium of the ecological system but also poses significant health hazards to humans. Microalgae emerged as a promising solution for achieving sustainability due to their ability to remediate contaminants and produce greener alternatives such as biofuels. This integrated approach provides an ambitious strategy to address global concerns pertaining to economic stability, environmental degradation, and the energy crisis. This study investigates the intricate defense mechanisms deployed by freshwater microalgae Chlorella minutissima in response to Cr (III) toxicity. The microalga achieved an impressive 92% removal efficiency with an IC50 value of 200 ppm, illustrating its extraordinary resilience towards chromium-induced stress. Furthermore, this research embarked on thorough explorations encompassing morphological, pigment-centric, and biochemical analyses, aimed at revealing the adaptive strategies associated with Cr (III) resilience, as well as the dynamics of carbon pool flow that contribute to enhanced lipid and extracellular polysaccharide (EPS) synthesis. The FAME profile of the biodiesel produced complies with the benchmark established by American and European fuel regulations, emphasizing its suitability as a high-quality vehicular fuel. Elevated levels of ROS, TBARS, and osmolytes (such as glycine-betaine), along with the increased activity of antioxidant enzymes (CAT, GR, and SOD), reveal the activation of robust defense mechanisms against oxidative stress caused by Cr (III). The finding of this investigation presents an effective framework for an algal-based biorefinery approach, integrating pollutant detoxification with the generation of vehicular-quality biodiesel and additional value-added compounds vital for achieving sustainability under the concept of a circular economy. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The IC<sub>50</sub> concentration of Cr (III) for <span class="html-italic">C. minutissima</span> after 96 h of exposure. (<b>B</b>) Growth curve comparison for <span class="html-italic">C. minutissima</span> under standard conditions and in the presence of Cr (III) at concentrations of 100 ppm and 200 ppm. (<b>C</b>) Changes in dry cell weight of <span class="html-italic">C. minutissima</span> over 12 days, with measurements taken every 72 h. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 indicated by * and **, respectively.</p>
Full article ">Figure 2
<p>(<b>A</b>) Removal efficiency of Cr (III) from the medium over time. (<b>B</b>) Time-course analysis of Cr (III) adsorbed on the cell surface of <span class="html-italic">C. minutissima</span>. (<b>C</b>) Quantification of Cr (III) accumulated within <span class="html-italic">C. minutissima</span> cells over 12 days, with measurements taken every 72 h. (<b>D</b>) Assessment of the bioconcentration factor for <span class="html-italic">C. minutissima</span> after 12 days of exposure to Cr (III) at concentrations of 100 ppm and 200 ppm. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05, 0.01 and 0.001 represented by *, **, and ***.</p>
Full article ">Figure 3
<p>XPS spectra of microalgal cells reveal Cr (III) in both oxide and hydroxide forms. (<b>A</b>) The first panel shows cells exposed to 100 ppm of Cr (III). (<b>B</b>) The second panel depicts cells treated with 200 ppm of Cr (III).</p>
Full article ">Figure 4
<p>(<b>A</b>) FESEM images (top panel) of <span class="html-italic">C. minutissima</span> cells exposed to 100 ppm and 200 ppm Cr (III), captured at a 2 μm scale and 5000× magnification, showing cellular morphology. The bottom panel displays EDX micrographs illustrating Cr (III) ion adsorption on cell surfaces under control conditions and at 100 ppm and 200 ppm Cr (III) concentrations. (<b>B</b>) Cell size measurements. (<b>C</b>) Zeta potential values. (<b>D</b>) FTIR spectra of <span class="html-italic">C. minutissima</span> cells under control conditions and after exposure to Cr (III) for 12 days. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 represented by *, respectively.</p>
Full article ">Figure 5
<p>Effects of different Cr (III) concentrations (100 ppm and 200 ppm) on (<b>A</b>) chlorophyll a levels; (<b>B</b>) chlorophyll b levels; (<b>C</b>) total chlorophyll content; (<b>D</b>) chlorophyll a/b ratio; (<b>E</b>) carotenoid content; and (<b>F</b>) PS II efficiency in <span class="html-italic">C. minutissima</span>. Measurements were taken every 72 h over a 12-day period. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 represented by *, and **.</p>
Full article ">Figure 6
<p>Changes in (<b>A</b>) carbonic anhydrase activity; (<b>B</b>) total ROS in terms of DCF fluorescence; (<b>C</b>) lipid peroxidation in terms of TBARS content; (<b>D</b>) total glycine betaine; (<b>E</b>) proline content; (<b>F</b>) catalase (CAT) activity; (<b>G</b>) ascorbate peroxidase (APX) activity; (<b>H</b>) glutathione reductase (GR) activity; and (<b>I</b>) superoxide dismutase (SOD) activity in <span class="html-italic">C. minutissima</span> cultivated in control and Cr (III)-spiked (100 ppm and 200 ppm) media for 12 days. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05, 0.01 and 0.001 represented by *, **, and ***.</p>
Full article ">Figure 7
<p>Time-based analysis of (<b>A</b>) lipid content; (<b>B</b>) carbohydrate content; and (<b>C</b>) protein content in <span class="html-italic">C. minutissima</span> biomass under control conditions and with Cr (III) exposure at 100 ppm and 200 ppm, over a 12-day incubation period. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 represented by * and **.</p>
Full article ">Figure 8
<p>(<b>A</b>) Bright field microscopy visuals (upper section) of <span class="html-italic">C. minutissima</span> after 12 days of growth (scale: 50 μm) and fluorescence microscopy visuals (lower section) of <span class="html-italic">C. minutissima</span> dyed with Nile red (scale: 50 μm). (<b>B</b>) Exemplary <sup>1</sup>H NMR spectra showcasing the overall lipid compositions of <span class="html-italic">C. minutissima</span> under standard conditions and after exposure to Cr (III) at concentrations of 100 ppm and 200 ppm. (<b>C</b>) Analysis of fold changes revealing shifts in various lipid categories. Within this context, the current study explores the integration of molecular mechanisms linked to Cr (III) tolerance, the rerouting of carbon metabolism triggered by Cr (III) towards lipid synthesis, and valuable by-products in the heavy metal-adaptive green microalga <span class="html-italic">C. minutissima</span>. Abbreviations: FA—fatty acid residue; TAG—triacylglyceride; PL—total phospholipid; MGDG—monogalactosyl diacylglycerol; PC—phosphatidylcholine; PE—phosphatidylethanolamine; PUFA—polyunsaturated fatty acid; omega-3-PUFA/ω3 PUFA—omega 3 polyunsaturated fatty acid. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 represented by * and **.</p>
Full article ">Figure 8 Cont.
<p>(<b>A</b>) Bright field microscopy visuals (upper section) of <span class="html-italic">C. minutissima</span> after 12 days of growth (scale: 50 μm) and fluorescence microscopy visuals (lower section) of <span class="html-italic">C. minutissima</span> dyed with Nile red (scale: 50 μm). (<b>B</b>) Exemplary <sup>1</sup>H NMR spectra showcasing the overall lipid compositions of <span class="html-italic">C. minutissima</span> under standard conditions and after exposure to Cr (III) at concentrations of 100 ppm and 200 ppm. (<b>C</b>) Analysis of fold changes revealing shifts in various lipid categories. Within this context, the current study explores the integration of molecular mechanisms linked to Cr (III) tolerance, the rerouting of carbon metabolism triggered by Cr (III) towards lipid synthesis, and valuable by-products in the heavy metal-adaptive green microalga <span class="html-italic">C. minutissima</span>. Abbreviations: FA—fatty acid residue; TAG—triacylglyceride; PL—total phospholipid; MGDG—monogalactosyl diacylglycerol; PC—phosphatidylcholine; PE—phosphatidylethanolamine; PUFA—polyunsaturated fatty acid; omega-3-PUFA/ω3 PUFA—omega 3 polyunsaturated fatty acid. The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 represented by * and **.</p>
Full article ">Figure 9
<p>FAME profile of control and Cr (III) spiked lyophilized biomass after 12 days of incubation.</p>
Full article ">Figure 10
<p>(<b>A</b>) Total soluble extracellular polysaccharide content (EPS). (<b>B</b>) Total carbohydrate content in soluble EPS. (<b>C</b>) FTIR spectra of lyophilized EPS powder from supernatant of control and Cr (III)-spiked microalgal cultures.The data denotes the mean ± S.D. from three independent replicates, with <span class="html-italic">p</span>-values &lt; 0.05 and 0.01 represented by * and **, respectively.</p>
Full article ">Figure 11
<p>(<b>A</b>) A schematic representation depicting the tolerance mechanism adopted by <span class="html-italic">C. minutissima</span> to survive under Cr (III) stress. Various steps involved in survival mechanism adopted by <span class="html-italic">C. minutissima</span> involves: (A). Interaction of Cr (III) ions with the functional groups present on algal cellular surface; (B) Impaired photosynthesis machinery due to Cr (III) stress; (C) Disruption of mitochondrial metabolic activities; (D) Generation of oxidative stress within the cell; (E) Activation of antioxidant machinery as defense mechanism; (F) Lipid peroxidation induced due to elevated levels of ROS; (G) Disruption of cellular membrane/membrane lipids; (H–I) in response <span class="html-italic">C. minutissima</span> shift carbon flux towards synthesis of energy reservoirs; (J) increased lipid production. (<b>B</b>) A conceptual illustration portraying an environmentally friendly biorefinery design that harnesses Cr (III) resistant microalgae for the purpose of environmental remediation while concurrently facilitating the production of biodiesel and an array of high-value substances like exopolysaccharides (EPS).</p>
Full article ">
15 pages, 6822 KiB  
Article
Application of Carboxymethyl Cellulose (CMC)-Coated Nanoscale Zero-Valent Iron in Chromium-Containing Soil Remediation
by Bo Zhang, Jiani Zhan, Jiaqi Fan, Bohong Zhu, Weili Shen, Shiwei Zhang, Weiting Li, Zhaohui Li and Fanjun Zeng
Clean Technol. 2024, 6(4), 1610-1624; https://doi.org/10.3390/cleantechnol6040078 - 11 Dec 2024
Viewed by 587
Abstract
Nanofine zero−valent iron (nZVI) is a new, eco−friendly material with strong reducing and adsorbent properties that can be used to clean up heavy metal−affected soils. Herein, nZVI encapsulated with carboxymethyl cellulose (CMC−nZVI) is synthesized via an aqueous−phase reduction technique and subsequently deployed to [...] Read more.
Nanofine zero−valent iron (nZVI) is a new, eco−friendly material with strong reducing and adsorbent properties that can be used to clean up heavy metal−affected soils. Herein, nZVI encapsulated with carboxymethyl cellulose (CMC−nZVI) is synthesized via an aqueous−phase reduction technique and subsequently deployed to evaluate its effectiveness in Cr(VI) soil remediation. The characterization analysis used SEM−EDS, XRD, XPS, and LSV to determine the relevant properties of the material. The results show that at an initial Cr(VI) concentration of 169.5 mg·kg−1, 93.2% of Cr(VI) was removed from the soil after 10 h of treatment with CMC−nZVI at pH 3.3. The kinetic analysis showed that CMC−nZVI had the maximum equilibrium adsorption capacity for removing Cr(VI) from soil at 105.3 mg·g−1. This followed a pseudo−second−order kinetic model. The study shows that CMC−nZVI converts Cr(VI) to Cr(III), which forms complexes with Fe(III) ions in the presence of hydroxide ions (OH) to form a highly stable compound that eventually adsorbs into the nanomaterial’s surface for efficient removal. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>a</b>) nZVI and (<b>b</b>) CMC−nZVI.</p>
Full article ">Figure 2
<p>XRD profiles of nZVI and CMC−nZVI.</p>
Full article ">Figure 3
<p>XPS pattern of (<b>a</b>) nZVI and (<b>b</b>) CMC−nZVI.</p>
Full article ">Figure 4
<p>LSV anodic polarization curves of nZVI and CMC−nZVI. (<b>b</b>) An enlarged view of the dashed box in (<b>a</b>).</p>
Full article ">Figure 5
<p>Images of antioxidant properties of different materials.</p>
Full article ">Figure 6
<p>(<b>a</b>) Pseudo−first−order kinetics model and (<b>b</b>) pseudo−second−order kinetics model.</p>
Full article ">Figure 7
<p>The adsorption isotherm of Cr (pH = 3.3, T = 25 °C) conforms to the Langmuir and Freundlich models.</p>
Full article ">Figure 8
<p>(<b>a</b>) SEM images of CMC−nZVI and (<b>b</b>) CMC−nZVI−Cr and (<b>c</b>,<b>d</b>) EDS images of CMC−nZVI−Cr. In (<b>d</b>), the blue color represents the C element, the red color represents the O element, the green color represents the Fe element, and the cyan color represents the Cr element.</p>
Full article ">Figure 9
<p>XRD images of CMC−nZVI and CMC−nZVI−Cr.</p>
Full article ">Figure 10
<p>E−pH diagram of the Cr−O<sub>2</sub>−H<sub>2</sub>O system.</p>
Full article ">Figure 11
<p>Removal mechanism for Cr(VI) by CMC−nZVI in soil.</p>
Full article ">
17 pages, 3482 KiB  
Article
Improving Lettuce Tolerance to Cadmium Stress: Insights from Raw vs. Cystamine-Modified Biochar
by Rongqi Chen, Xi Duan, Ruoxuan Xu and Tao Zhao
Horticulturae 2024, 10(12), 1323; https://doi.org/10.3390/horticulturae10121323 - 11 Dec 2024
Viewed by 291
Abstract
Understanding the interactions among biochar, plants, soils, and microbial communities is essential for developing effective and eco-friendly soil remediation strategies. This study investigates the role of cystamine-modified biochar (Cys-BC) in alleviating cadmium (Cd) toxicity in lettuce, comparing its effects to those of raw [...] Read more.
Understanding the interactions among biochar, plants, soils, and microbial communities is essential for developing effective and eco-friendly soil remediation strategies. This study investigates the role of cystamine-modified biochar (Cys-BC) in alleviating cadmium (Cd) toxicity in lettuce, comparing its effects to those of raw biochar. Lettuce plants were exposed to Cd stress (1–5 mg kg−1), and the effects of Cys-BC were assessed by measuring plant biomass, photosynthetic efficiency, antioxidant activity, Cd bioavailability, and soil microbial diversity. Cys-BC significantly enhanced plant biomass, with increases in above-ground growth (40.54–44.95%) and root biomass (37.54–47.44%) compared to Cd-stressed controls. Photosynthetic parameters improved by up to 91.02% for chlorophyll-a content and 37.93% for the net photosynthetic rate. Cys-BC mitigated oxidative stress, increasing antioxidant activities by 73.83% to 99.39%. Additionally, Cys-BC reduced available Cd levels in the soil, primarily through enhanced cation exchange rather than changes in pH. Plant responses to Cd stress included increased glutathione reductase activity and elevated cysteine levels, which further contributed to Cd passivation. Microbial diversity in the soil increased, particularly among sulfur- and nitrogen-cycling bacteria such as Deltaproteobacteria and Nitrospira, suggesting their role in mitigating Cd stress. These findings highlight the potential of Cys-BC as an effective agent for the remediation of Cd-contaminated soils. Full article
(This article belongs to the Special Issue Microbial Interaction with Horticulture Plant Growth and Development)
Show Figures

Figure 1

Figure 1
<p>Effects of different biochar treatments on oxidative stress markers content in lettuce: (<b>a</b>) malondialdehyde, (<b>b</b>) hydrogen peroxide, (<b>c</b>) glutathione, and (<b>d</b>) cysteine. Lowercase (a–d) letters are used to denote differential rankings.</p>
Full article ">Figure 2
<p>Effects of different biochar treatments on the activity of antioxidant enzymes of lettuce: (<b>a</b>) superoxide dismutase, (<b>b</b>) peroxidase, (<b>c</b>) catalase, and (<b>d</b>) reductase. Lowercase (a–d) letters are used to denote differential rankings.</p>
Full article ">Figure 3
<p>The linear regression models correlating Cd residues in lettuce with DTPA-extractable Cd in soil. (<b>a</b>) shoot Cd and (<b>b</b>) root Cd under raw BC treatments; (<b>c</b>) shoot Cd and (<b>d</b>) root Cd under Cys-BC treatments.</p>
Full article ">Figure 4
<p>Effects of different biochar treatments on soil properties: (<b>a</b>) proportion of various Cd fractions, (<b>b</b>) DPTA available Cd content, (<b>c</b>) soil pH, and (<b>d</b>) soil CEC.</p>
Full article ">Figure 5
<p>The <span class="html-italic">t</span>-test for the variability in the abundance of soil bacteria and fungi in the different biochar treatments, (<b>a</b>) differences in bacteria of raw BC treatments at the phylum level, (<b>b</b>) differences in bacteria of Cys-BC treatments at the phylum level, (<b>c</b>) differences in fungi at the phylum level, (<b>d</b>) differences in bacteria of raw BC treatments at the program level, (<b>e</b>) differences in bacteria of Cys-BC treatments at the program level, (<b>f</b>) differences in fungi at the program level, (<b>g</b>) differences in bacteria of raw BC treatments at the order level, (<b>h</b>) differences in bacteria in Cys-BC treatments at the order level, (<b>i</b>) differences in fungi at the order level. Analyses of variance levels of significance (LS): * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.01, ns: not significant.</p>
Full article ">Figure 6
<p>Canonical correspondence analysis (CCA) of bacteria (<b>a</b>,<b>b</b>) fungi in soil sample groups.</p>
Full article ">
18 pages, 4051 KiB  
Article
Photosynthetic Efficiency of Plants as an Indicator of Tolerance to Petroleum-Contaminated Soils
by Piotr Dąbrowski, Ilona Małuszyńska, Marcin J. Małuszyński, Bogumiła Pawluśkiewicz, Tomasz Gnatowski, Aneta H. Baczewska-Dąbrowska and Hazem M. Kalaji
Sustainability 2024, 16(24), 10811; https://doi.org/10.3390/su162410811 - 10 Dec 2024
Viewed by 467
Abstract
Significant efforts have been made to develop environmentally friendly remediation methods to restore petroleum-damaged ecosystems. One such approach is cultivating plant species that exhibit high resistance to contamination. This study aimed to assess the impact of petroleum-derived soil pollutants on the photosynthetic performance [...] Read more.
Significant efforts have been made to develop environmentally friendly remediation methods to restore petroleum-damaged ecosystems. One such approach is cultivating plant species that exhibit high resistance to contamination. This study aimed to assess the impact of petroleum-derived soil pollutants on the photosynthetic performance of selected plant species used in green infrastructure development. A pot experiment was conducted using both contaminated and uncontaminated soils to grow six plant species under controlled conditions. Biometric parameters and chlorophyll a fluorescence measurements were taken, followed by statistical analyses to compare plant responses under stress and control conditions. This study is the first to simultaneously analyze PF, DF, and MR820 signals in plant species exposed to petroleum contamination stress. The results demonstrated that petroleum exposure reduced the activity of both PSII and PSI, likely due to increased nonradiative energy dissipation in PSII antenna chlorophylls, decreased antenna size, and/or damage to the photosynthetic apparatus. Additionally, petroleum contamination affected the electron transport chain efficiency, limiting electron flow between PSII and PSI. The most resistant species to petroleum-induced stress were Lolium perenne, Poa pratensis, and Trifolium repens. Full article
(This article belongs to the Section Pollution Prevention, Mitigation and Sustainability)
Show Figures

Figure 1

Figure 1
<p>Weather conditions during the experiment period.</p>
Full article ">Figure 2
<p>Induction curves of chlorophyll <span class="html-italic">a</span> fluorescence of investigated plants under control soil and petroleum-contaminated soil. <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH), relative units, n = 8.</p>
Full article ">Figure 3
<p>Delayed fluorescence induction curves of investigated plants under control soil and petroleum-contaminated soil. <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH), relative units, n = 8.</p>
Full article ">Figure 4
<p>Modulated light reflection at 820 nm of investigated plants under control soil and petroleum-contaminated soil. <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH), relative units, n = 8.</p>
Full article ">Figure 5
<p>Principal component analysis for control experiment (<b>a</b>) and petroleum contamination soil (<b>b</b>). <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH).</p>
Full article ">Figure 6
<p>T-Student test results of all measured parameters of tested plant species. <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH). The ns refer to non-significant differences, and asterisks *, **, ***, **** refer to different significance levels represented by <span class="html-italic">p</span>-values lower or equal to 0.05, 0.01, 0.001 and 0.0001 respectively.</p>
Full article ">Figure 7
<p>Boxplot results of the parameters Area (<b>a</b>), MRmin (<b>b</b>), and I2 (<b>c</b>) indicating statistically significant (or no-significant) differences in the fluorescence variables enable the impact of the contamination treatment on the state of the analyzed species. <span class="html-italic">Dactylis glomerata</span> L. var Amba (DGA), <span class="html-italic">Lolium perenne</span> L. var. Maja (LPM) and Nira (LPN), <span class="html-italic">Poa pretensis</span> L. var. Appalachian (PPA), <span class="html-italic">Phleum pretense</span> L. var. Egida (PPE), <span class="html-italic">Trifolium repens</span> L. var Grass. Huia (TRH). The blue box and circles regard control experiment, whereas the green box and triangles indicate the Petroleum contamination experiment.</p>
Full article ">
16 pages, 5345 KiB  
Article
Enhancing Pb Adsorption on Crushed Microplastics: Insights into the Environmental Remediation
by Sen Li, Lu Cao, Qiyuan Liu, Shuting Sui, Jiayin Bian, Xizeng Zhao and Yun Gao
Water 2024, 16(23), 3541; https://doi.org/10.3390/w16233541 - 9 Dec 2024
Viewed by 643
Abstract
This study investigates the pollution characteristics and environmental risks of crushed microplastics (MPs) generated during plastic recycling, emphasizing their adsorption capacity for heavy metals, particularly lead (Pb). SEM-EDS analysis revealed that crushed MPs exhibit significantly higher adsorption capacity than primary MPs due to [...] Read more.
This study investigates the pollution characteristics and environmental risks of crushed microplastics (MPs) generated during plastic recycling, emphasizing their adsorption capacity for heavy metals, particularly lead (Pb). SEM-EDS analysis revealed that crushed MPs exhibit significantly higher adsorption capacity than primary MPs due to their larger surface area and more available adsorption sites, including oxygen-containing functional groups. The adsorption behavior of MPs was influenced by key factors such as MP size, MP quantity, pH, salinity, and biofilm formation. Smaller MPs demonstrated higher adsorption efficiency, while elevated pH enhanced Pb adsorption. Conversely, increased salinity reduced adsorption due to competition for adsorption sites. Increasing MP concentrations improved Pb removal efficiency, but higher MP quantities led to a decrease in maximum adsorption capacity, demonstrating a trade-off between removal efficiency and adsorption capacity. Isothermal adsorption experiments revealed that Pb adsorption on MPs follows a multi-layer mechanism, best characterized by the Freundlich model. The adsorption capacity increased nonlinearly with Pb concentration and stabilized as surface sites became saturated. The formation of biofilms on MPs further enhanced their adsorption capacity by providing additional functional groups and facilitating multi-layer adsorption, increasing ecological risks. Adsorption kinetics were best described by pseudo-second-order and intra-particle diffusion models, indicating chemical adsorption and boundary layer diffusion as dominant mechanisms. Magnetic Fe3O4 nanoparticles demonstrated a high recovery efficiency of 99.3% for MPs, highlighting their potential for environmental remediation. However, the presence of adsorbed Pb slightly reduced recovery performance, emphasizing the need to optimize recovery conditions for maximum efficiency. These findings underscore the dual threat posed by crushed MPs: their capacity to adsorb and concentrate harmful substances, increasing ecological toxicity, and the challenges associated with their recovery. This research provides critical insights into mitigating MP pollution and developing effective recovery strategies under realistic environmental conditions. Full article
(This article belongs to the Topic Microplastics Pollution)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SEM image of primary MPs (magnification: 10.00KX); (<b>b</b>) SEM image and corresponding elemental composition (EDS) bar chart of crushed MPs; (magnification: 10.00K X) (<b>c</b>) SEM image and corresponding elemental composition (EDS) bar chart of Pb-adsorbed MPs (magnification: 10.00K X).</p>
Full article ">Figure 2
<p>The adsorption of Pb on primary MPs and crushed MPs with time.</p>
Full article ">Figure 3
<p>Adsorption curves of crushed MPs with different particle sizes in Pb solution. Experimental conditions: [adsorbents] = 1 g/L; [Pb] = 30 mg/L.</p>
Full article ">Figure 4
<p>Adsorption characteristics of crushed MPs in Pb solutions: adsorption curves (<b>a</b>), adsorption kinetics (<b>b</b>), and intra-particle diffusion model (<b>c</b>). Experimental conditions: [adsorbents] = 1 g/L; [MP size] = 0.2 mm.</p>
Full article ">Figure 5
<p>Adsorption isotherm of heavy metal Pb solution on MPs. Experimental conditions: [adsorbent] = (0.1 g/L, 0.2 g/L, 0.3 g/L); [MP size] = 0.2 mm.</p>
Full article ">Figure 6
<p>Effect of the pH (<b>a</b>) and salinity (<b>b</b>) on the adsorption of Pb on crushed PVC. Experimental conditions: [adsorbents] = 0.3 g/L; [MP size] = 0.2 mm.</p>
Full article ">Figure 7
<p>Adsorption curves of crushed MPs with different concentrations in Pb solution. Experimental conditions: [MP size] = 0.2 mm; [Pb] = 30 mg/L.</p>
Full article ">Figure 8
<p>Biofilm formation on the surface of MPs (magnification: 350X).</p>
Full article ">Figure 9
<p>Effect of biofilm-coated MPs and non-coated MPs on Pb adsorption in crushed PVC. Experimental conditions: [adsorbent] = 1 g/L; [Pb] = 30 mg/L.</p>
Full article ">Figure 10
<p>The effect of Fe<sub>3</sub>O<sub>4</sub> mass on the recovery efficiency of MPs.</p>
Full article ">Figure 11
<p>The effect of adsorption capacity of MPs on their recovery efficiency by Fe<sub>3</sub>O<sub>4</sub> nanoparticles in Pb-contaminated wastewater.</p>
Full article ">
Back to TopTop