[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,418)

Search Parameters:
Keywords = dye wastewater

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 4510 KiB  
Article
Effect of Calcination Temperature on the Photocatalytic Activity of Precipitated ZnO Nanoparticles for the Degradation of Rhodamine B Under Different Light Sources
by Amira Saidani, Reguia Boudraa, Karim Fendi, Lamia Benouadah, Abderrahim Benabbas, Atmane Djermoune, Stefano Salvestrini, Jean-Claude Bollinger, Abdulmajeed Abdullah Alayyaf and Lotfi Mouni
Water 2025, 17(1), 32; https://doi.org/10.3390/w17010032 - 26 Dec 2024
Viewed by 234
Abstract
This research provides valuable insights into the application of ZnO nanoparticles in photocatalytic wastewater treatment. Process optimization was carried out by determining the ratio of the surface area to the energy band gap (S/E) in the photocatalysis rate under different sources of light [...] Read more.
This research provides valuable insights into the application of ZnO nanoparticles in photocatalytic wastewater treatment. Process optimization was carried out by determining the ratio of the surface area to the energy band gap (S/E) in the photocatalysis rate under different sources of light (UV light, visible light, sunlight). The nanoparticles were synthesized using the precipitation technique, and the calcination process was carried out within a temperature range of 400 to 700 °C. The structural, morphological, and optical properties of materials were investigated using X-ray powder diffraction (XRD), scanning electron microscopy (SEM), UV-Vis diffuse reflectance (UV-Vis DRS), Raman spectroscopies, and Fourier transform infrared (FTIR) spectroscopies. The study demonstrates that calcination temperature significantly influences the photocatalytic activity of ZnO nanoparticles by altering their size, surface properties, shape, and optical behavior. Optimal decomposition efficiencies of Rhodamine B were achieved at 400 °C, with yields of 24%, 92%, and 91% under visible, UV, and sunlight irradiation, respectively. Additionally, the surface area decreased from 12.556 to 8.445 m2/g, the band gap narrowed slightly from 3.153 to 3.125 eV, and crystal growth increased from 0.223 to 0.506 µm as the calcination temperature rose. The photocatalytic properties of ZnO nanoparticles were assessed to determine their efficiency in decomposing Rhodamine B dye under operational parameters, including pollutant concentration (C0), sample amount, pH level, and reaction time. The sample exhibited the best breakdown rates with C0 = 5 mg/L, solid-to-liquid ratio (S/L) = 50 mg/L, pH = 7, and reaction time = 1 h. Additionally, we combined two oxidation processes, namely H2O2 and photocatalytic oxidation processes, which significantly improved the Rhodamine B removal efficiency, where 100% of RhB was degraded after 60 min and 100 µL of H2O2. Full article
(This article belongs to the Special Issue Advanced Biotechnologies for Water and Wastewater Treatment)
Show Figures

Figure 1

Figure 1
<p>XRD spectra of ZnO calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 2
<p>SEM image for different samples of ZnO nanoparticles calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 3
<p>FTIR spectra of ZnO nanoparticles calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 4
<p>Nitrogen adsorption–desorption isotherms for synthesized ZnO samples calcined at (<b>a</b>) 400 °C, (<b>b</b>) 500 °C, (<b>c</b>) 600 °C, and (<b>d</b>) at 700 °C.</p>
Full article ">Figure 5
<p>Raman spectrum of ZnO nanoparticles calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 6
<p>The optical properties and Tauc plots for determining the band gap energy of the photocatalyst calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 7
<p>UV-visible absorption spectra of ZnO nanoparticles calcined at 400 °C, 500 °C, 600 °C, and 700 °C.</p>
Full article ">Figure 8
<p>Adsorption capacity of ZnO calcined at different temperature for Rhodamine B at different calcination temperatures (pH = 6, S/L = 50 mg/100 mL, C<sub>0</sub> = 5 mg/L, time = 60 min).</p>
Full article ">Figure 9
<p>Degradation of RhB under visible (<b>a</b>), UV (<b>b</b>), and sunlight (<b>c</b>) irradiation (pH = 6, S/L = 50 mg/100 mL, C<sub>0</sub> = 5 mg/L, time = 60 min).</p>
Full article ">Figure 10
<p>Degradation mechanism of RhB by ZnO.</p>
Full article ">Figure 11
<p>(<b>a</b>) Mass effect of ZnO degradation of RhB (pH = 6, C<sub>0</sub> = 5 mg/L), (<b>b</b>) effect of initial concentration of RhB on degradation rate (pH = 6, S/L = 50 mg/100 mL), (<b>c</b>) pH Effect for the degradation of RhB by ZnO under sunlight (S/L= 50 mg/100 mL, C<sub>0</sub> = 6 mg/L), and (<b>d</b>) H<sub>2</sub>O<sub>2</sub> effect on the degradation of RhB (pH = 7.4, S/L = 50 mg/100 mL, C<sub>0</sub> = 6 mg/L).</p>
Full article ">Figure 12
<p>Scavenger effect on photodegradation of RhB by ZnO under sunlight, pH = 6, S/L = 50 mg/100 mL, C<sub>0</sub> = 3 mg/L and [Scavenger] = 1 × 10<sup>−5</sup> M.</p>
Full article ">
11 pages, 3376 KiB  
Article
Ultra-Fast Removal of CBB from Wastewater by Imidazolium Ionic Liquids-Modified Nano-Silica
by Mengyue Zhang, Fan Yang, Nan Wang, Jifu Du, Juntao Yan, Ya Sun, Manman Zhang and Long Zhao
Molecules 2025, 30(1), 24; https://doi.org/10.3390/molecules30010024 - 25 Dec 2024
Viewed by 26
Abstract
The efficient removal of dyes is of significant importance for environmental purification and human health. In this study, a novel material (Si-MPTS-IL) has been synthesized by the immobilization of imidazole ionic liquids (ILs) onto nano-silica using the radiation grafting technique. The adsorption performance [...] Read more.
The efficient removal of dyes is of significant importance for environmental purification and human health. In this study, a novel material (Si-MPTS-IL) has been synthesized by the immobilization of imidazole ionic liquids (ILs) onto nano-silica using the radiation grafting technique. The adsorption performance of Si-MPTS-IL for Coomassie Brilliant Blue (CBB) removal is studied by a series of static adsorption experiments. It is found that Si-MPTS-IL has ultra-fast adsorption kinetics, reaching equilibrium within 2 min. The adsorption process for CBB conforms to the Langmuir model. In addition, Si-MPTS-IL exhibits a negligible impact on the adsorption efficiency of CBB with the increase in salt concentration. After six cycles of adsorption–desorption, the adsorption efficiency of Si-MPTS-IL remained above 80%, indicating excellent regenerative properties and a promising candidate for the treatment of wastewater containing CBB. A study of the mechanism indicates that the CBB capture by Si-MPTS-IL can be attributed to the synergistic effects of electrostatic interactions and pore filling. Full article
Show Figures

Figure 1

Figure 1
<p>Synthesis route of Si-MPTS-IL.</p>
Full article ">Figure 2
<p>TEM (<b>a</b>), N<sub>2</sub> adsorption–desorption isotherms (<b>b</b>) and pore width distribution (<b>c</b>) of F-Si, Si-MPTS and Si-MPTS-IL.</p>
Full article ">Figure 3
<p>Zeta potential of Si-MPTS-IL (<b>a</b>); Effect of pH (<b>b</b>) and contact time (<b>c</b>) on CBB uptake; (<b>d</b>) 3D fluorescence spectrogram during adsorption on Si-MPTS-IL.</p>
Full article ">Figure 4
<p>Effect of initial concentration (<b>a</b>), temperature (<b>b</b>), salt concentration (<b>c</b>) and cycle number (<b>d</b>) on the adsorption of CBB by Si-MPTS-IL.</p>
Full article ">Figure 5
<p>FT-IR spectra of Si-MPTS-IL before and after adsorption.</p>
Full article ">
14 pages, 5070 KiB  
Article
Magnetically Assembled Electrode Incorporating Self-Powered Tourmaline Composite Particles: Exploiting Waste Energy in Electrochemical Wastewater Treatment
by Bo Zhang, Dan Shao, Yaru Wang, Hao Xu and Haojie Song
Catalysts 2025, 15(1), 2; https://doi.org/10.3390/catal15010002 - 24 Dec 2024
Viewed by 34
Abstract
A magnetically assembled electrode (MAE) is a modular electrode format in electrochemical oxidation wastewater treatment. MAE utilizes magnetic forces to attract the magnetic catalytic auxiliary electrodes (AEs) on the main electrode (ME), which has the advantages of high efficiency and flexible adjustability. However, [...] Read more.
A magnetically assembled electrode (MAE) is a modular electrode format in electrochemical oxidation wastewater treatment. MAE utilizes magnetic forces to attract the magnetic catalytic auxiliary electrodes (AEs) on the main electrode (ME), which has the advantages of high efficiency and flexible adjustability. However, the issue of the insufficient polarization of the AEs leaves the potential of this electrode underutilized. In this study, natural tourmaline (Tml) particles with pyroelectric and piezoelectric properties were utilized to solve the above issue by harvesting and converting the waste energy (i.e., the joule heating energy and the bubble striking mechanical energy) from the electrolysis environment into additional electrical energy applied on the AEs. Different contents of Tml particles were composited with Fe3O4/Sb-SnO2 particles as novel AEs, and the structure–activity relationship of the novel MAE was investigated by various electrochemical measurements and orthogonal tests of dye wastewater treatment. The results showed that Tml could effectively enhance all electrochemical properties of the electrode. The optimal dye removal rate was obtained by loading the AEs with 0.2 g·cm−2 when the Tml content was 4.5 wt%. The interaction of current density and Tml content had a significant effect on the COD removal rate, and the mineralization capacity of the electrode was significantly enhanced. The findings of this study have unveiled the potential application of minerals and energy conversion materials in the realm of electrochemical oxidation wastewater treatment. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Preparation processes of different AEs particles in this study and their material characterization results: (<b>a</b>) SEM image and particle size distribution of Tml. (<b>b</b>) SEM images and particle size distribution of SnO<sub>2</sub>(0T). (<b>c</b>) EDS elemental content distribution of SnO<sub>2</sub>(0T). SEM images and particle size distribution of (<b>d</b>) SnO<sub>2</sub>(4.5%T) and (<b>e</b>) SnO<sub>2</sub>(16%T). (<b>f</b>) Tml polarization curves; (<b>g</b>) XRD images of the three AEs. (<b>h</b>) Schematic diagram of the distribution of different ratios of tourmaline doping.</p>
Full article ">Figure 2
<p>Electrochemical characterization of 2D Ti/Sb-SnO<sub>2</sub> and each group of MAE: (<b>a</b>) Double-layer capacitance value (C<sub>dl</sub>). (<b>b</b>) Voltametric charge (Q*) obtained at different potential scan rates and the corresponding q<sub>T</sub>. (<b>c</b>) CV curves (potential range: 0~2.5 V (vs. SCE), scan rate: 0.01 V·s<sup>−1</sup>). (<b>d</b>) Tafel plots of LSV curves. (<b>e</b>,<b>f</b>) Nyquist plots (equilibrium potential: 0 V and 2 V (vs. SCE), frequency range: 0.1~10<sup>5</sup> Hz). (<b>g</b>) Comprehensive comparison radar plots of key electrochemical performance metrics.</p>
Full article ">Figure 3
<p>One-factor experiments on the degradation of ARG (250 mL, 200 mg·L<sup>−1</sup>) by four electrodes composed of Ti/Sb-SnO<sub>2</sub> for 90 min under four experimental conditions: (<b>a</b>) ARG removal rate versus time; (<b>b</b>) COD of ARG solution after 90 min of degradation.</p>
Full article ">Figure 4
<p>Results of orthogonal test analysis based on ARG removal rate after 30 min of degradation: (<b>a</b>) Distribution of contributions of significant single and interaction factors to experimental results. (<b>b</b>) Significant single factor main effects at each level. (<b>c</b>,<b>d</b>) Space curved surface plot of the effect of different interaction factors on ARG removal rate.</p>
Full article ">Figure 5
<p>Results of orthogonal test analysis based on COD removal rate after 90 min of degradation: (<b>a</b>) Distribution of contributions of significant single and interaction factors to experiment results. (<b>b</b>) Significant single factor main effects at each level. (<b>c</b>,<b>d</b>) Space curved surface plot of the effect of different interaction factors on COD removal rate.</p>
Full article ">Scheme 1
<p>Structure of the magnetically assembled electrode (MAE) and the novel tourmaline composite auxiliary electrodes (AEs) particles in this study and the schematic diagram of the waste energy conversion of tourmaline in electrolysis.</p>
Full article ">
14 pages, 4072 KiB  
Article
High-Performance Photocatalytic Degradation—A ZnO Nanocomposite Co-Doped with Gd: A Systematic Study
by Aeshah Alasmari, Nadi Mlihan Alresheedi, Mohammed A. Alzahrani, Fahad M. Aldosari, Mostafa Ghasemi, Atef Ismail and Abdelaziz M. Aboraia
Catalysts 2024, 14(12), 946; https://doi.org/10.3390/catal14120946 - 20 Dec 2024
Viewed by 446
Abstract
This research aims to analyze the improvement in the photocatalytic properties of ZnO nanoparticles by incorporating Gd. In order to understand the influence of incorporating Gd into the ZnO matrix, the photocatalytic activity of the material is compared at various Gd concentrations. Different [...] Read more.
This research aims to analyze the improvement in the photocatalytic properties of ZnO nanoparticles by incorporating Gd. In order to understand the influence of incorporating Gd into the ZnO matrix, the photocatalytic activity of the material is compared at various Gd concentrations. Different doping concentrations of Gd ranging from 0 to 0.075 are incorporated into ZnO and the synthesized ZnO-Gd nanocomposites are investigated using structural, morphological, and optical analyses using XRD, SEM, and UV-vis spectroscopy, respectively. The photocatalytic performance of the synthesized ZnO-Gd nanocomposites is determined via the degradation of organic contaminants under visible light. Regarding the latter, the results suggest that photocatalytic efficiency increases with increasing Gd doping levels up to an optimal doping concentration. The enhancement of the photocatalytic performance of Gd-doped ZnO is explained, along with the mechanism related to the availability of new pathways for charge carrier recombination. Among all of them, the 0.075 Gd-doped ZnO catalyst exhibits the highest photocatalytic activity which degrades 89% of MB dye after being irradiated with UV light for 120 min. However, pure ZnO degrades only 40% of MB dye within the same testing conditions. In closing, this work confirms the applicability of Gd-doped ZnO nanocomposites as photocatalysts in cleaning up the environment and in wastewater treatment. Full article
(This article belongs to the Special Issue Design and Application of Combined Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>a</b>) ZnO, (<b>b</b>) 0.025 Gd-doped ZnO, (<b>c</b>) 0.05 Gd-doped ZnO, and (<b>d</b>) 0.075 Gd-doped ZnO.</p>
Full article ">Figure 2
<p>FTIR spectroscopy of (<b>a</b>) ZnO, 0.05 Gd-doped ZnO, and 0.075 Gd-doped ZnO and (<b>b</b>) selected peaks at around 417, 419, 443 cm<sup>−1</sup> for ZnO, 0.05 Gd-doped ZnO, and 0.075 Gd-doped ZnO.</p>
Full article ">Figure 3
<p>UV-vis spectroscopy (<b>a</b>,<b>b</b>) energy gap calculated of ZnO, 0.025 Gd-doped ZnO, 0.05 Gd-doped ZnO, and 0.075 Gd-doped ZnO.</p>
Full article ">Figure 4
<p>SEM photos of (<b>a</b>) ZnO, (<b>b</b>) 0.025 Gd-doped ZnO, (<b>c</b>) 0.05 Gd-doped ZnO, (<b>d</b>) 0.075 Gd-doped ZnO.</p>
Full article ">Figure 5
<p>Photocatalytic degradation of (<b>a</b>) ZnO, (<b>b</b>) 0.025 Gd-doped ZnO, (<b>c</b>) 0.05 Gd-doped ZnO, and (<b>d</b>) 0.075 Gd-doped ZnO.</p>
Full article ">Figure 6
<p>Removal efficiency of ZnO, 0.025 Gd-doped ZnO, 0.05 Gd-doped ZnO, and 0.075 Gd-doped ZnO.</p>
Full article ">Figure 7
<p>–ln C<sub>t</sub>/C<sub>0</sub> vs. time of ZnO, 0.025 Gd-doped ZnO, 0.05 Gd-doped ZnO, and 0.075 Gd-doped ZnO.</p>
Full article ">Figure 8
<p>EIS of (<b>a</b>) ZnO (in set equivalent circuit), (<b>b</b>) 0.025 Gd-doped ZnO (in set equivalent circuit), (<b>c</b>) 0.05 Gd-doped ZnO (in set equivalent circuit), and (<b>d</b>) 0.075 Gd-doped ZnO (in set equivalent circuit).</p>
Full article ">Figure 9
<p>Mott-Schotky plot of (<b>a</b>) ZnO, (<b>b</b>) 0.025 Gd-doped ZnO, (<b>c</b>) 0.05 Gd-doped ZnO, and (<b>d</b>) 0.075 Gd-doped ZnO.</p>
Full article ">Figure 10
<p>Photocatalytic reaction mechanism.</p>
Full article ">Figure 11
<p>Recycle test of 0.075 Gd-doped ZnO.</p>
Full article ">
18 pages, 10877 KiB  
Proceeding Paper
Development of an Operationally Efficient and Cost-Effective System for Removal of Dye from Wastewater Using Novel Adsorbent
by Niraj S. Topare, Sunita Raut-Jadhav and Anish Khan
Eng. Proc. 2024, 76(1), 102; https://doi.org/10.3390/engproc2024076102 - 19 Dec 2024
Viewed by 211
Abstract
An important contributor to environmental degradation is the industrial revolution, which has occurred in developed and developing nations. The present investigation aimed to tackle the escalating apprehensions regarding the discharge of various types of dyes from the paint, textile, and dyeing sectors. This [...] Read more.
An important contributor to environmental degradation is the industrial revolution, which has occurred in developed and developing nations. The present investigation aimed to tackle the escalating apprehensions regarding the discharge of various types of dyes from the paint, textile, and dyeing sectors. This research focuses on the adsorption performance of a newly developed system that uses cotton pod shell powder (CPSP) as a novel adsorbent to remove dye industry wastewater. The system has been designed, manufactured, and tested to be operationally efficient and cost-effective. The CPSP is a new adsorbent with desirable properties such as favorable functional groups and porosity, and analysis of its functional groups and porous nature was carried out using FTIR and SEM. The experimental data from the developed system showed that inlet dye concentration (50, 100, and 150 ppm), bed height (10, 20, and 30 cm), and flow rate (10, 15, and 20 mL/min) significantly affect the adsorption of dye industry wastewater by CPSP. Breakthrough curves were shown to be flow rate and bed depth dependent, according to the data. Significant experimentation was conducted on the developed system, and under optimized conditions. It was shown that the breakthrough point was affected by both bed height and flow rate. Evidence suggested that decreasing flow rate and concentration and raising bed height led to improved breakthrough and exhaustion times. At a concentration of 100 ppm and a flow rate of 15 mL/min, a bed depth of 20 cm was found to have the highest absorption capacity. Adam-Bohart, bed depth service time, and Yoon-Nelson models were utilized to examine the adsorption data. The results revealed that the developed system is effective, and the data obtained in this work can provide optimum operating conditions, suggesting its scalability to an industrial level for dye removal from wastewater by adsorption using CPSP as a novel adsorbent. Full article
Show Figures

Figure 1

Figure 1
<p>Various dye removal methods.</p>
Full article ">Figure 2
<p>Processing of CPSP.</p>
Full article ">Figure 3
<p>SEM image of CPSP.</p>
Full article ">Figure 4
<p>FTIR for CPSP.</p>
Full article ">Figure 5
<p>Fixed-bed column adsorption investigation experimental setup.</p>
Full article ">Figure 6
<p>Breakthrough curve for CPSP at an initial dye concentration.</p>
Full article ">Figure 7
<p>Breakthrough curve for dye on CPSP at different flow rates.</p>
Full article ">Figure 8
<p>Breakthrough curve for dye on CPSP at different bed heights.</p>
Full article ">Figure 9
<p>Bohart-Adams model to CPSP for different bed heights.</p>
Full article ">Figure 10
<p>Bohart-Adams model to CPSP for different flow rates.</p>
Full article ">Figure 11
<p>Bohart-Adams model to CPSP for different initial dye concentrations.</p>
Full article ">Figure 12
<p>BDST model to CPSP for different bed heights.</p>
Full article ">Figure 13
<p>BDST model to CPSP for different flow rates.</p>
Full article ">Figure 14
<p>BDST model to CPSP for different initial dye concentrations.</p>
Full article ">Figure 15
<p>Yoon-Nelson model to CPSP for different bed heights.</p>
Full article ">Figure 16
<p>Yoon-Nelson model to CPSP for different flow rates.</p>
Full article ">Figure 17
<p>Yoon-Nelson model to CPSP for different initial dye concentrations.</p>
Full article ">Figure 18
<p>Desorption of dye industry wastewater adsorbed on CPSP.</p>
Full article ">
25 pages, 8553 KiB  
Article
Employment of Fe3O4/Fe2TiO5/TiO2 Composite Made Using Ilmenite for Elimination of Methylene Blue
by Himasha Gunathilaka and Charitha Thambiliyagodage
ChemEngineering 2024, 8(6), 130; https://doi.org/10.3390/chemengineering8060130 - 18 Dec 2024
Viewed by 507
Abstract
A novel material was created from natural ilmenite sand, and methylene blue (MB) was used to test the material’s capacity to remove colors from wastewater. The material was synthesized by neutralizing the acid leachate obtained by Ilmenite sand digestion, followed by drying at [...] Read more.
A novel material was created from natural ilmenite sand, and methylene blue (MB) was used to test the material’s capacity to remove colors from wastewater. The material was synthesized by neutralizing the acid leachate obtained by Ilmenite sand digestion, followed by drying at 180 °C. It was characterized by XRD, Raman, TEM, SEM, XPS, XRF, and BET techniques. The crystal nature of the composite is Fe3O4/Fe2TiO5/TiO2. The surface area, average pore size and total pore volume of the composite are 292.18 m2/g, 1.53 nm, and 0.202 cc/g, respectively. At pH 10, 10 mg/L MB, and 10 mg of the material resulted in a maximum adsorption capacity of 24.573 mg/g. Using 5 mg/L increments, the dye concentration was adjusted between 10 and 25 mg/L, yielding equilibrium adsorption capacities of 24.573, 31.012, 41.443, and 52.259 mg/g with 10, 15, 20, and 25 mg/L, respectively. The greatest adsorbent capacity of 24.573 mg/g was achieved with 10 mg of the adsorbent and 10 mg/L MB. The adsorbent dosage ranged from 10, 25, 45, 65, and 100 mg. MB was adsorbed via pseudo-second-order kinetics with an adsorption capacity of 24.863 mg/g. The intraparticle diffusion model showed that MB adsorption occurs in three stages, with intra-particle diffusion constants of 1.50, 2.71, 3.38, and 4.41 g/mg min1/2. Adsorption followed the Langmuir isotherm model. The obtained thermodynamic parameters ΔG, ΔH, and ΔS were −27.5521 kJ/mol at 298 K, 2.571 kJ/mol, and 0.101 kJ/mol, respectively. Regeneration studies of the adsorbent were carried out for five cycles, indicating some activity loss after each cycle. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 2
<p>Raman spectrum of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 3
<p>High-resolution XPS spectra of (<b>a</b>) Ti 2p, (<b>b</b>) Fe 2p, (<b>c</b>) O 2p, (<b>d</b>) C 2p, of composite (<b>e</b>) Survey spectrum.</p>
Full article ">Figure 3 Cont.
<p>High-resolution XPS spectra of (<b>a</b>) Ti 2p, (<b>b</b>) Fe 2p, (<b>c</b>) O 2p, (<b>d</b>) C 2p, of composite (<b>e</b>) Survey spectrum.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image; (<b>b</b>) TEM image; (<b>c</b>,<b>d</b>) HR-TEM images of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 5
<p>(<b>a</b>) Nitrogen adsorption–desorption isotherms of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> nanocomposite. (<b>b</b>) Pore size distribution of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 6
<p>(<b>a</b>) Effect of pH on adsorption of MB on Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub>. (<b>b</b>) Determination of the point of zero charge of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub>.</p>
Full article ">Figure 7
<p>Variation in adsorption capacity of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> on (<b>a</b>) different concentrations of MB and (<b>b</b>) different dosages of adsorbent.</p>
Full article ">Figure 8
<p>Effect of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> dosage on MB adsorption capacity.</p>
Full article ">Figure 9
<p>(<b>a</b>) Pseudo-first-order and (<b>b</b>) pseudo-second-order kinetics model for adsorption of MB to Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 10
<p>Intraparticle diffusion model for adsorption of MB to Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 11
<p>Boyd diffusion model for adsorption of MB to Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 12
<p>Elovich model for adsorption of MB to Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 13
<p>(<b>a</b>) Langmuir, (<b>b</b>) Freundlich, (<b>c</b>) Temkin, and (<b>d</b>) Dubinin adsorption isotherm models of MB adsorption onto Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 13 Cont.
<p>(<b>a</b>) Langmuir, (<b>b</b>) Freundlich, (<b>c</b>) Temkin, and (<b>d</b>) Dubinin adsorption isotherm models of MB adsorption onto Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 14
<p>Van’t Hoff plots of MB adsorption onto Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub> composite.</p>
Full article ">Figure 15
<p>(<b>a</b>) Pseudo-second-order kinetic model plotted for the adsorption of MB at four different temperatures and (<b>b</b>) Arrhenius plots constructed for the adsorption of Fe<sub>3</sub>O<sub>4</sub>/Fe<sub>2</sub>TiO<sub>5</sub>/TiO<sub>2</sub>.</p>
Full article ">Figure 16
<p>(<b>a</b>) C/C<sub>0</sub> vs. time (<b>b</b>) equilibrium adsorption capacity of five adsorption–desorption cycles.</p>
Full article ">
29 pages, 12851 KiB  
Article
Nanosilver–Biopolymer–Silica Composites: Preparation, and Structural and Adsorption Analysis with Evaluation of Antimicrobial Properties
by Magdalena Blachnio, Malgorzata Zienkiewicz-Strzalka, Jolanta Kutkowska and Anna Derylo-Marczewska
Int. J. Mol. Sci. 2024, 25(24), 13548; https://doi.org/10.3390/ijms252413548 - 18 Dec 2024
Viewed by 357
Abstract
In this article, we report on the research on the synthesis of composites based on a porous, highly ordered silica material modified by a metallic nanophase and chitosan biofilm. Due to the ordered pore system of the SBA-15 silica, this material proved to [...] Read more.
In this article, we report on the research on the synthesis of composites based on a porous, highly ordered silica material modified by a metallic nanophase and chitosan biofilm. Due to the ordered pore system of the SBA-15 silica, this material proved to be a good carrier for both the biologically active nanophase (highly dispersed silver nanoparticles, AgNPs) and the adsorption active phase (chitosan). The antimicrobial susceptibility was determined against Gram-positive Staphylococcus aureus ATCC 25923, Gram-negative bacterial strains (Escherichia coli ATCC 25922, Klebsiella pneumoniae ATCC 700603, and Pseudomonas aeruginosa ATCC 27853), and yeast Candida albicans ATCC 90028. The zones of microbial growth inhibition correlated with the content of silver nanoparticles deposited in the composites and were the largest for C. albicans (14–21 mm) and S. aureus (12–17 mm). The suitability of the composites for the purification of water and wastewater from anionic pollutants was evaluated based on kinetic and equilibrium adsorption studies for the dye Acid Red 88. The composite with the highest amount of the chitosan component showed the greatest adsorption capacity (am) of 0.57 mmol/g and the most effective kinetics with a rate constant (log k) and half-time (t0.5) of −0.21 and 1.62 min, respectively. Due to their great practical importance, AgNP–chitosan–silica composites can aspire to be classified as functional materials combining the environmental problem with microbiological activity. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the action of silver nanoparticles on bacterial cells using the example of the <span class="html-italic">E. coli</span> bacteria model.</p>
Full article ">Figure 2
<p>The process of silver ion reduction controlled by UV–Vis technique: (<b>A</b>) changes in the concentration of silver nanoparticles in the process of reduction of diamminesilver(I) ions; and (<b>B</b>) UV–Vis spectra of silver nanoparticle solutions recorded after specific time intervals from the establishment of equilibrium. Each spectrum in <a href="#ijms-25-13548-f002" class="html-fig">Figure 2</a>B was recorded in the wavelength range of 300 to 900 nm.</p>
Full article ">Figure 3
<p>UV–Vis spectra for pure SBA-15 and composites with different contents of silver nanoparticles.</p>
Full article ">Figure 4
<p>(<b>A</b>) Nitrogen adsorption–desorption isotherms at 77 K for the analyzed samples, (<b>B</b>) porosity distributions calculated with the BJH theory from the adsorption, and (<b>C</b>) desorption branches of the isotherms as a function of pore size (in a differential form dV/dD, where V and D are pore volume and diameter, respectively).</p>
Full article ">Figure 4 Cont.
<p>(<b>A</b>) Nitrogen adsorption–desorption isotherms at 77 K for the analyzed samples, (<b>B</b>) porosity distributions calculated with the BJH theory from the adsorption, and (<b>C</b>) desorption branches of the isotherms as a function of pore size (in a differential form dV/dD, where V and D are pore volume and diameter, respectively).</p>
Full article ">Figure 5
<p>Dependences of surface charge density of the composites AgChS1–AgChS3 on solution pH.</p>
Full article ">Figure 6
<p>The powder XRD patterns in the range of small diffraction angles for AgChS1–AgChS3 samples.</p>
Full article ">Figure 7
<p>XRD images of the biopolymer–nanosilver composites AgChS1–AgChS3 and experimental comparison curves of chitosan and silica components.</p>
Full article ">Figure 8
<p>Parameterization of the first XRD Ag(111) signal of the tested materials AgChS1 (<b>A</b>), AgChS2 (<b>B</b>), and AgChS3 (<b>C</b>). Inset tables show calculations of the crystallite size relative to the plane perpendicular to the (111), (200), (220), and (311) directions together with the average silver crystallite size D (nm).</p>
Full article ">Figure 8 Cont.
<p>Parameterization of the first XRD Ag(111) signal of the tested materials AgChS1 (<b>A</b>), AgChS2 (<b>B</b>), and AgChS3 (<b>C</b>). Inset tables show calculations of the crystallite size relative to the plane perpendicular to the (111), (200), (220), and (311) directions together with the average silver crystallite size D (nm).</p>
Full article ">Figure 9
<p>Transmission electron micrographs (TEMs) for the AgNP–chitosan–silica composites AgChS1 (<b>A</b>,<b>B</b>), AgChS2 (<b>C</b>,<b>D</b>), and AgChS3 (<b>E</b>,<b>F</b>).</p>
Full article ">Figure 10
<p>Microbial growth inhibition of: <span class="html-italic">S. aureus</span> ATCC 25923, <span class="html-italic">E. coli</span> ATCC 25922, <span class="html-italic">K. pneumoniae</span> ATCC 700603, <span class="html-italic">P. aeruginosa</span> ATCC 27853, and <span class="html-italic">C. albicans</span> ATCC 90028 by tested materials: AgChS1–AgChS3, ChSBA, AgNPs, ampicillin (AMP), and amphotericin B (AMPH).</p>
Full article ">Figure 11
<p>AFM images of Gram-positive <span class="html-italic">Staphylococcus aureus</span> untreated (control sample (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and Staphylococcus aureus exposed to AgNP–chitosan–silica composite (AgChS3 (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>)) (<b>A</b>,<b>B</b>) topography as 2D view and 3D image of the upper surface of bacterial strains (<b>C</b>,<b>D</b>), enlargement of the bacterial envelope area of the control bacteria (<b>E</b>) and the bacteria exposed to the composite material (<b>F</b>), and surface topography of bacteria before (<b>G</b>) and after contact with the material (<b>H</b>).</p>
Full article ">Figure 11 Cont.
<p>AFM images of Gram-positive <span class="html-italic">Staphylococcus aureus</span> untreated (control sample (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and Staphylococcus aureus exposed to AgNP–chitosan–silica composite (AgChS3 (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>)) (<b>A</b>,<b>B</b>) topography as 2D view and 3D image of the upper surface of bacterial strains (<b>C</b>,<b>D</b>), enlargement of the bacterial envelope area of the control bacteria (<b>E</b>) and the bacteria exposed to the composite material (<b>F</b>), and surface topography of bacteria before (<b>G</b>) and after contact with the material (<b>H</b>).</p>
Full article ">Figure 12
<p>AFM images of Gram-negative <span class="html-italic">Escherichia coli</span> untreated (control sample (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and bacteria exposed to AgChS3 (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>)) (<b>A</b>,<b>B</b>) topography as 2D view and 3D image of the upper surface of bacterial strains (<b>C</b>,<b>D</b>), enlargement of the bacterial envelope area of the control bacteria (<b>E</b>) and the bacteria exposed to the composite material (<b>F</b>), and surface topography of bacteria before (<b>G</b>) and after contact with the composite material (<b>H</b>).</p>
Full article ">Figure 12 Cont.
<p>AFM images of Gram-negative <span class="html-italic">Escherichia coli</span> untreated (control sample (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and bacteria exposed to AgChS3 (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>)) (<b>A</b>,<b>B</b>) topography as 2D view and 3D image of the upper surface of bacterial strains (<b>C</b>,<b>D</b>), enlargement of the bacterial envelope area of the control bacteria (<b>E</b>) and the bacteria exposed to the composite material (<b>F</b>), and surface topography of bacteria before (<b>G</b>) and after contact with the composite material (<b>H</b>).</p>
Full article ">Figure 13
<p>(<b>A</b>) Adsorption isotherms of Acid Red 88 on the AgNP–chitosan–silica composites and on the composite ChSBA (as a control material, inset) as a dependence of adsorbed dye amount on equilibrium concentration c<sub>eq</sub>. (<b>B</b>) Dependence of adsorption capacity a<sub>m</sub> on the nitrogen content in the composites.</p>
Full article ">Figure 14
<p>Comparison of Acid Red 88 adsorption kinetics on the AgNP–chitosan–silica composites at coordinates: relative concentration~time (<b>A</b>,<b>C</b>); relative concentration~square root of time (<b>B</b>,<b>D</b>). The lines correspond to the fitted m-exponential equation (<b>A</b>,<b>B</b>) and the fractal-like SOE equation (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 15
<p>Dependence of standard deviations of relative concentration SD(c)/c<sub>0</sub> on the number of exponential terms in the multi-exponential equation (<b>A</b>); the relationship between adsorption kinetics at fixed values of the process progress and N content in the composites (<b>B</b>); and distribution of half-time t<sub>0.5i</sub> (<b>C</b>) and rate coefficient k<sub>i</sub> (<b>D</b>) for dye adsorption on the composites.</p>
Full article ">Figure 16
<p>Scheme of the synthesis of stabilized silver nanoparticles solution.</p>
Full article ">
22 pages, 4715 KiB  
Article
A Hybrid Photo-Catalytic Approach Utilizing Oleic Acid-Capped ZnO Nanoparticles for the Treatment of Wastewater Containing Reactive Dyes
by Zakia H. Alhashem, Ashraf H. Farha, Shrouq H. Aleithan, Shehab A. Mansour and Maha A. Tony
Catalysts 2024, 14(12), 934; https://doi.org/10.3390/catal14120934 - 18 Dec 2024
Viewed by 314
Abstract
In pursuit of overcoming Fenton oxidation limitations in wastewater treatment, an introduction of a heterogeneous photocatalyst was developed. In this regard, the current work introduces ZnO nanocrystals that were successfully prepared via a thermal decomposition technique and then capped with oleic acid (OA). [...] Read more.
In pursuit of overcoming Fenton oxidation limitations in wastewater treatment, an introduction of a heterogeneous photocatalyst was developed. In this regard, the current work introduces ZnO nanocrystals that were successfully prepared via a thermal decomposition technique and then capped with oleic acid (OA). The synthesized ZnO-OA and the pristine ZnO were characterized by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and field emission scanning electron microscopy (FE-SEM). Then, the study introduces the application of such materials in advanced oxidation processes, i.e., a Fenton reaction to treat dye-containing wastewater. Synthetic wastewater that was prepared using Reactive Blue 4 (RB4) was used as a simulated textile wastewater effluent. Fenton’s oxidation was applied, and the system parameters were assessed using the modified Fenton’s system. The synthesized samples of ZnO were characterized by a recognized wurtzite hexagonal structure. The surface modification of ZnO with oleic acid (OA) resulted in an increase in crystallite size, lattice parameters, and cell volume. These modifications were linked to the efficient capping of ZnO nanoparticles by OA, which further improved the dispersion of the nanoparticles, as demonstrated through SEM imaging. The optimum conditions of ZnO- and ZnO-OA-synthesized modified Fenton composites showed 400 mg/L and 40 mg/L for H2O2 and the catalyst, respectively, at pH 3.0, and within 90 min under UV irradiation the maximal dye oxidation reached 93%. The catalytic performance at its optimal circumstances was in accordance with a pseudo-second-order kinetics model for both ZnO-OA- and the pristine ZnO-based Fenton’s systems. The thermodynamic parameters, including the enthalpy (ΔH′), the entropy (ΔS′), and Gibbs free energy (ΔG′) of activations, were also checked, and their values settled that both ZnO and ZnO-OA Fenton systems are non-spontaneous in nature. Furthermore, the reaction signified for processing at a low energy barrier condition (10.38 and 31.38 kJ/mol for ZnO-OA- and the pristine ZnO-based Fenton reactions, respectively). Full article
(This article belongs to the Special Issue Design and Synthesis of Nanostructured Catalysts, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns with Rietveld analysis according to ZnO standard data for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals.</p>
Full article ">Figure 2
<p>The Williamson–Hall plot (4sinθ vs. β<math display="inline"><semantics> <mrow> <mi>cos</mi> <mi mathvariant="sans-serif">θ</mi> </mrow> </semantics></math>) of the ZnO and ZnO-OA nanocrystals.</p>
Full article ">Figure 3
<p>FE-SEM micrographs of the synthesized ZnO nanocrystals (<b>a</b>) 7000× magnification and 10 μm scale and ZnO-OA nanocrystals (<b>b</b>) 12,000× magnification and 10 μm scale.</p>
Full article ">Figure 4
<p>Effect of reaction time on different oxidation systems.</p>
Full article ">Figure 5
<p>The Effect of RB4 dye loading on the oxidation rate (operating conditions at pH 3.0; H<sub>2</sub>O<sub>2</sub> 400 mg/L and catalyst 40 mg/L) for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton’s systems.</p>
Full article ">Figure 6
<p>The effect of catalyst concentration on RB4 dye oxidation at pH 3.0 and H<sub>2</sub>O<sub>2</sub> 400 mg/L for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton systems.</p>
Full article ">Figure 7
<p>The effect of H<sub>2</sub>O<sub>2</sub> concentration on RB4 dye oxidation at pH 3.0 and the catalyst at 40 mg/L for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton’s systems.</p>
Full article ">Figure 8
<p>The effect of pH on RB4 dye oxidation at hydrogen peroxide 400 mg/L and catalyst 40 mg/L for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton’s systems.</p>
Full article ">Figure 9
<p>The effect of temperature on RB4 dye oxidation at pH 3.0, hydrogen peroxide at 400 mg/L, and the catalyst at 40 mg/L for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton’s systems.</p>
Full article ">Figure 10
<p>Plots of the kinetic models on the oxidation of RB4 dye with Fenton reagent at different temperatures for zero-order ((<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals based Fenton’s systems), pseudo-first-order ((<b>c</b>) ZnO and (<b>d</b>) ZnO-OA based Fenton’s systems), and pseudo-second-order ((<b>e</b>) ZnO- and (<b>f</b>) ZnO-OA-based Fenton’s systems).</p>
Full article ">Figure 11
<p>The Arrhenius plot of the second-order kinetic constants for (<b>a</b>) ZnO and (<b>b</b>) ZnO-OA nanocrystals-based Fenton’s systems.</p>
Full article ">Figure 12
<p>Schematic diagram for the capping of as-synthesized ZnO nanocrystals with oleic acid.</p>
Full article ">Figure 13
<p>Graphical representation of the treatment steps and experimental set-up.</p>
Full article ">
15 pages, 9112 KiB  
Article
Efficient Dye Contaminant Elimination and Simultaneous Electricity Production via a Carbon Quantum Dots/TiO2 Photocatalytic Fuel Cell
by Zixuan Feng, Xuechen Li, Yueying Lv and Jie He
Crystals 2024, 14(12), 1083; https://doi.org/10.3390/cryst14121083 - 16 Dec 2024
Viewed by 381
Abstract
Conventional wastewater treatment methods do not fully utilize the energy in wastewater. This study uses a photocatalytic fuel cell (PFC) to remove dye impurities and generate electricity with that energy. Pt serves as the PFC’s cathode, while the carbon quantum dots (CQDs)/anatase TiO [...] Read more.
Conventional wastewater treatment methods do not fully utilize the energy in wastewater. This study uses a photocatalytic fuel cell (PFC) to remove dye impurities and generate electricity with that energy. Pt serves as the PFC’s cathode, while the carbon quantum dots (CQDs)/anatase TiO2 (A-TiO2) serve as its photoanode. The visible light absorption range of A-TiO2 can be increased by combining CQDs with A-TiO2. The composite of CQD and A-TiO2 broadens the absorption edge from 364 nm to 538 nm. TiO2’s different crystal structures and particle sizes impact the PFC’s power generation and dye contaminant removal. The 30 min photodegradation rate of methylene blue by the 20 nm A-TiO2 was 97.3%, higher than that of the 5 nm A-TiO2 (75%), 100 nm A-TiO2 (92.1%), and A-TiO2 (93%). The photocurrent density of the 20 nm A-TiO2 can reach 4.41 mA/cm2, exceeding that of R-TiO2 (0.64 mA/cm2), 5 nm A-TiO2 (1.97 mA/cm2), and 100 nm A-TiO2 (3.58 mA/cm2). The photodegradative and electrochemical test results show that the 20 nm A-TiO2 delivers a better degradation and electrochemical performance than other samples. When the 20 nm A-TiO2 was used in the PFC photoanode, the photocurrent density, open-circuit voltage, and maximum power density of the PFC were found to be 0.6 mA/cm2, 0.41 V, and 0.1 mW/cm2, respectively. The PFC prepared in this study shows a good level of performance compared to recent similar systems. Full article
(This article belongs to the Special Issue Synthesis and Properties of Photocatalysts)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD of R-TiO<sub>2</sub>, A-TiO<sub>2</sub>, and 10% CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) A-TiO<sub>2</sub> and (<b>b</b>) 10% CQDs/A-TiO<sub>2</sub>; TEM images of (<b>c</b>) TiO<sub>2</sub> and (<b>d</b>) 10% CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 3
<p>The A-TiO<sub>2</sub> and 10% CQDs/A-TiO<sub>2</sub> XPS spectra: (<b>a</b>) survey spectra, (<b>b</b>) O 1s, (<b>c</b>) C 1s, and (<b>d</b>) Ti 2p.</p>
Full article ">Figure 4
<p>(<b>a</b>) Typical UV-VIS absorption spectra and (<b>b</b>) the Tauc plot of R-TiO<sub>2</sub>, A-TiO<sub>2</sub>, and 10% CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 5
<p>FTIR spectra of 10% CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>Photodegradation curves (<b>a</b>) and pseudo-first-order rate kinetics curves of (<b>b</b>) R-TiO<sub>2</sub>, 5 nm A-TiO<sub>2</sub>, 20 nm A-TiO<sub>2</sub>, 100 nm A-TiO<sub>2</sub>, and 10% CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 7
<p>Photocurrent density profiles of (<b>a</b>) A-TiO<sub>2</sub> with different particle sizes and R-TiO<sub>2</sub> and (<b>b</b>) different ratios of CQDs/A-TiO<sub>2</sub>.</p>
Full article ">Figure 8
<p>PFC: (<b>a</b>) photocurrent density curve; (<b>b</b>) open-circuit voltage curve; (<b>c</b>) polarization curve; (<b>d</b>) power density curve.</p>
Full article ">Figure 9
<p>Photodegradation curves of PFC at different voltages.</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 441
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 ">
13 pages, 4035 KiB  
Communication
Use of Laccase Enzymes as Bio-Receptors for the Organic Dye Methylene Blue in a Surface Plasmon Resonance Biosensor
by Araceli Sánchez-Álvarez, Gabriela Elizabeth Quintanilla-Villanueva, Osvaldo Rodríguez-Quiroz, Melissa Marlene Rodríguez-Delgado, Juan Francisco Villarreal-Chiu, Analía Sicardi-Segade and Donato Luna-Moreno
Sensors 2024, 24(24), 8008; https://doi.org/10.3390/s24248008 - 15 Dec 2024
Viewed by 623
Abstract
Methylene blue is a cationic organic dye commonly found in wastewater, groundwater, and surface water due to industrial discharge into the environment. This emerging pollutant is notably persistent and can pose risks to both human health and the environment. In this study, we [...] Read more.
Methylene blue is a cationic organic dye commonly found in wastewater, groundwater, and surface water due to industrial discharge into the environment. This emerging pollutant is notably persistent and can pose risks to both human health and the environment. In this study, we developed a Surface Plasmon Resonance Biosensor employing a BK7 prism coated with 3 nm chromium and 50 nm of gold in the Kretschmann configuration, specifically for the detection of methylene blue. For the first time, laccases immobilized on a gold surface were utilized as bio-receptors for this organic dye. The enzyme was immobilized using carbodiimide bonds with EDC/NHS crosslinkers, allowing for the analysis of samples with minimal preparation. The method demonstrated validation with a limit of detection (LOD) of 4.61 mg L−1 and a limit of quantification (LOQ) of 15.37 mg L−1, a working range of 0–100 mg L−1, and an R2 value of 0.9614 during real-time analysis. A rainwater sample spiked with methylene blue yielded a recovery rate of 122.46 ± 4.41%. The biosensor maintained a stable signal over 17 cycles and remained effective for 30 days at room temperature. Full article
(This article belongs to the Section Biosensors)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of methylene blue.</p>
Full article ">Figure 2
<p>Possible recognition process and first step of degradation of methylene blue by laccases.</p>
Full article ">Figure 3
<p>Immobilization process of laccases on the thin chromium–gold film chip. In the step 1, alkanethiols are added to the thin gold surface. In step 2, the EDC is added, forming an unstable intermediate. In step 3, the NHS is added, creating a sulfo-NHS ester. In step 4, NHS is replaced by the laccase through an amide bond.</p>
Full article ">Figure 4
<p>Assembly of the prism and the chip on the SPR equipment: (<b>a</b>) assembly of the prism, the chip with a thin gold film with the immobilized laccases, the prism and other components. (<b>b</b>) Set up of the prism, sample cell, chip with immobilized laccases and the other on the SPR equipment.</p>
Full article ">Figure 5
<p>Reflectance spectra obtained by angular sweep.</p>
Full article ">Figure 6
<p>Immobilization process of laccass from <span class="html-italic">Rhus vernicifera</span> in real-time by SPR.</p>
Full article ">Figure 7
<p>FTIR analysis of different stages of laccase immobilization.</p>
Full article ">Figure 8
<p>(<b>a</b>) SPR analysis of stocks with different concentrations of methylene blue. (<b>b</b>) Calibration curve and equation of a straight line.</p>
Full article ">Figure 9
<p>Comparison of the intensity of reflectance of solutions of methylene blue at day 1 and day 30.</p>
Full article ">
22 pages, 6321 KiB  
Review
A Review of Innovative Cucurbituril-Based Photocatalysts for Dye Degradation
by Mosab Kaseem
Catalysts 2024, 14(12), 917; https://doi.org/10.3390/catal14120917 (registering DOI) - 12 Dec 2024
Viewed by 1495
Abstract
This review explores the advancements in photocatalysis facilitated by cucurbiturils (CBs), specifically focusing on CB[5], CB[6], CB[7], and CB[8]. Cucurbiturils have gained prominence due to their exceptional ability to enhance photocatalytic reactions through mechanisms such as improved charge separation, high adsorption capacities, and [...] Read more.
This review explores the advancements in photocatalysis facilitated by cucurbiturils (CBs), specifically focusing on CB[5], CB[6], CB[7], and CB[8]. Cucurbiturils have gained prominence due to their exceptional ability to enhance photocatalytic reactions through mechanisms such as improved charge separation, high adsorption capacities, and the generation of reactive oxygen species. The review summarizes recent research on the use of CBs in various photocatalytic applications, including dye degradation, pollutant removal, and wastewater treatment. Studies highlight CB[5]’s utility in dye removal and the creation of efficient nanocomposites for improved degradation rates. CB[6] is noted for its high adsorption capacities and photocatalytic efficiency in both adsorption and degradation processes. CB[7] shows promise in adsorbing and degrading toxic dyes and enhancing fluorescence in biomedical applications, while CB[8] leads to significant improvements in photocatalytic activity and stability. The review also discusses the synthesis, properties, and functionalization of cucurbiturils to maximize their photocatalytic potential. Future research directions include the optimization of cucurbituril-based composites, the exploration of new application areas, and scaling up their use for practical environmental and industrial applications. This comprehensive review provides insights into the current capabilities of cucurbituril-based photocatalysts and identifies key areas for future development in sustainable photocatalytic technologies. Full article
(This article belongs to the Special Issue Green Chemistry and Catalysis)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Acid-catalyzed synthesis of cucurbituril (CB[n]) homologues through the condensation of glycoluril (1) and formaldehyde. (<b>b</b>) Various representations of the CB[7] structure, highlighting its chemical framework, three-dimensional geometry, and dimensional parameters [<a href="#B34-catalysts-14-00917" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>Chemical structures of cucurbiturils with double cavities and their schematic representations [<a href="#B17-catalysts-14-00917" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) the 3D view of the complex resulted from the hybridization (W<sub>6</sub>O<sub>19</sub>)<sup>2−</sup> and (Me<sub>10</sub>CB[5]). (<b>b</b>) Decolorization rates of RhB solutions with different materials [<a href="#B67-catalysts-14-00917" class="html-bibr">67</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image of pure CdS nanoparticles showing large aggregates. (<b>b</b>) SEM image of CdS/CB[5] composite with smaller, dispersed particles due to CB[5]. (<b>c</b>) Photocatalytic degradation of Methylene Blue (MB) with different CdS/CB[5] proportions under visible light, showing enhanced degradation with more CB[5]. (<b>d</b>) Comparison of MB degradation between pure CdS and CdS/CB[5], demonstrating the higher efficiency of the CdS/CB[5] composite [<a href="#B69-catalysts-14-00917" class="html-bibr">69</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Structural assembly of the CB[6]-POM composite, illustrating the hydrogen bonding interactions between the {Ni-CB[6]}n chain and the POM chain. (<b>b</b>) Decolorization rates of methyl orange (MO) under different reaction conditions; MO/photocatalyst/H<sub>2</sub>O<sub>2</sub>, MO/photocatalyst, and MO/H<sub>2</sub>O<sub>2</sub>. (<b>c</b>) Proposed dual-pathway mechanism for the photodegradation of MO involving superoxide radicals (O<sub>2</sub><sup>•−</sup>) and hydroxyl radicals (<sup>•</sup>OH) [<a href="#B74-catalysts-14-00917" class="html-bibr">74</a>].</p>
Full article ">Figure 6
<p>Regeneration and real-water performance of CB/MMCNT adsorbent: (<b>a</b>–<b>c</b>) adsorption capacity retention after seven cycles for MB, MG, and GV; (<b>d</b>) removal efficiency in real water samples, showing over 92% effectiveness [<a href="#B78-catalysts-14-00917" class="html-bibr">78</a>].</p>
Full article ">Figure 7
<p>Optimized structures of the 2CB[7]@TfT complex, showing stabilization with and without magnesium cations. Magnesium cations act as “lids”, enhancing the complex’s stability and fluorescence performance [<a href="#B83-catalysts-14-00917" class="html-bibr">83</a>].</p>
Full article ">Figure 8
<p>Formation of 2D Polypseudorotaxanes with Cucurbit[8]uril (CB[8]). CB[8] binds to naphthol-modified porphyrin (TPP-Np) and viologen derivatives (DMV) through charge-transfer interactions, forming a stable, single-layer 2D network [<a href="#B92-catalysts-14-00917" class="html-bibr">92</a>].</p>
Full article ">
18 pages, 6580 KiB  
Article
Evaluation of Almond Shell Activated Carbon for Dye (Methylene Blue and Malachite Green) Removal by Experimental and Simulation Studies
by Adrián Rial, Catarina Helena Pimentel, Diego Gómez-Díaz, María Sonia Freire and Julia González-Álvarez
Materials 2024, 17(24), 6077; https://doi.org/10.3390/ma17246077 - 12 Dec 2024
Viewed by 406
Abstract
The present work analyzes the behavior of an activated carbon fabricated from almond shells for the removal of cationic dyes (methylene blue, MB, and malachite green, MG) by adsorption from aqueous solutions. The carbonized precursor was activated with KOH at a 1:2 ( [...] Read more.
The present work analyzes the behavior of an activated carbon fabricated from almond shells for the removal of cationic dyes (methylene blue, MB, and malachite green, MG) by adsorption from aqueous solutions. The carbonized precursor was activated with KOH at a 1:2 (w/w) ratio with the objective of increasing both the surface area and the pore volume. Both non-activated and activated carbon were characterized in different aspects of interest in dye adsorption studies (surface structure, point of zero charge, specific surface area, and pore size distribution). The effect of the dye’s initial concentration and adsorbent dosage on dye removal efficiency and carbon adsorption capacity was studied. Adsorption kinetics were analyzed under different experimental conditions, and different models were assayed to determine the adsorption mechanism. Dye adsorption in the adsorbent surface could be considered the rate-limiting step. Different adsorption equilibrium models were evaluated to fit the experimental data. This adsorbent allowed us to reach high Langmuir adsorption capacity for both dyes (MB: 341 mg·g−1, MG: 364 mg·g−1 at 25 °C and 0.5 g·L−1). Moreover, kinetic and equilibrium adsorption data have been used to simulate breakthrough curves in a packed-bed column using different conditions (bed length, liquid flowrate, and dye initial concentration). The simulation results showed that almond shell activated carbon is a suitable adsorbent for methylene blue and malachite green removal from wastewater. Full article
Show Figures

Figure 1

Figure 1
<p>Point of zero charge (pH<sub>pzc</sub>) determination for almond shell carbons.</p>
Full article ">Figure 2
<p>Activated carbon surface images obtained with scanning electron microscopy (SEM) before (<b>a</b>) and after methylene blue (<b>b</b>) and malachite green (<b>c</b>) adsorption (C<sub>0</sub> = 250 mg·L<sup>−1</sup>, adsorbent dosage = 0.5 g·L<sup>−1</sup>, T = 25 °C, t = 1440 min, natural pH).</p>
Full article ">Figure 3
<p>FTIR spectra of carbons before and after dye adsorption.</p>
Full article ">Figure 4
<p>Nitrogen adsorption and desorption isotherms at 77 K for almond shell non-activated and activated carbons.</p>
Full article ">Figure 5
<p>Carbon dioxide adsorption isotherms at 273 K for almond shell non-activated and activated carbons.</p>
Full article ">Figure 6
<p>Pore size distribution of the almond shell activated carbon estimated using NLDFT model.</p>
Full article ">Figure 7
<p>Influence of adsorbent dosage on adsorption percentage (columns) and capacity (symbols + lines) of MB (blue) and MG (green). T = 25 °C, t = 24 h. Solid line: 50 mg·L<sup>−1</sup>; dashed line: 250 mg·L<sup>−1</sup>; dotted line: 500 mg·L<sup>−1</sup>.</p>
Full article ">Figure 8
<p>Influence of dye initial concentration on adsorption percentage (columns) and capacity (symbols + lines) of MB (blue) and MG (green). T = 25 °C, t = 24 h. Solid line: 0.5 g·L<sup>−1</sup>; dashed line: 0.75 g·L<sup>−1</sup>; dotted line: 1 g·L<sup>−1</sup>.</p>
Full article ">Figure 9
<p>Adsorption kinetics for the adsorption of MB (<b>a</b>) and MG (<b>b</b>) adsorption by almond shell activated carbon using different initial dye concentrations at 25 °C, natural pH, and an adsorbent dosage of 0.5 g·L<sup>−1</sup>. Solid line corresponds to the fitting to the pseudo-second-order model.</p>
Full article ">Figure 10
<p>Experimental data and adsorption isotherm models corresponding to (<b>a</b>) MB and (<b>b</b>) MG adsorption by almond shell activated carbon at 25 °C, natural pH, and adsorbent dosage of 0.5 g·L<sup>−1</sup>.</p>
Full article ">Figure 11
<p>Proposal of the interactions involved in the adsorption mechanism for MB and MG adsorption.</p>
Full article ">Figure 12
<p>Simulated breakthrough curves for dyes adsorption by activated carbon: Red curves, MB adsorption in coconut shell carbon [<a href="#B34-materials-17-06077" class="html-bibr">34</a>]. Blue curves, MB adsorption in almond shell carbon. Green curves, MG adsorption in almond shell carbon. Continuous lines: 50 mg·L<sup>−1</sup>, initial dye concentration. Dashed line: 200 mg·L<sup>−1</sup>, initial dye concentration. Flowrate, Q<sub>L</sub> = 0.01 mL·min<sup>−1</sup>. Mass of carbon, m<sub>b</sub> = 12.1 g.</p>
Full article ">
22 pages, 1457 KiB  
Review
Environmental Impacts and Biological Technologies Toward Sustainable Treatment of Textile Dyeing Wastewater: A Review
by Yuqi Liu, Junsheng Chen, Dianrong Duan, Ziyang Zhang, Chang Liu, Wei Cai and Ziwen Zhao
Sustainability 2024, 16(24), 10867; https://doi.org/10.3390/su162410867 - 11 Dec 2024
Viewed by 1003
Abstract
Textile, printing, and dyeing industries in China are expanding annually, resulting in the discharge of significant volumes of wastewater. These effluents have complex compositions and contain diverse pollutants that pose severe hazards to aquatic systems, ecological environments, and nearby flora, fauna, and human [...] Read more.
Textile, printing, and dyeing industries in China are expanding annually, resulting in the discharge of significant volumes of wastewater. These effluents have complex compositions and contain diverse pollutants that pose severe hazards to aquatic systems, ecological environments, and nearby flora, fauna, and human populations. The inadequate or rudimentary treatment of these effluents can cause substantial environmental damage. Current technologies for treating textile dyeing wastewater (TDW) include physical, chemical, and biological methods, with biological treatment being noted for its low cost and environmental sustainability. In the realm of biotechnological treatment, microorganisms, such as bacteria, fungi, and algae, exhibit significant potential. This review highlights the urgent need for effective treatment of textile dyeing wastewater (TDW), which poses severe environmental and health risks. It provides a comparative analysis of physical, chemical, and biological treatment methods, with a focus on the unique advantages of biological approaches, such as biodegradation and biosorption, for sustainable wastewater management. Key findings include recent advancements in microbial applications, challenges in scaling up, and integration into existing treatment systems. This review aims to guide future research and practical applications in achieving eco-friendly and cost-effective solutions for TDW remediation. Full article
(This article belongs to the Special Issue Advances in Technologies for Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Key components of textile and dyeing wastewater and their health impacts.</p>
Full article ">Figure 2
<p>The general process of algae treatment of printing and dyeing wastewater.</p>
Full article ">
15 pages, 3951 KiB  
Article
Research on the Adsorption Mechanism and Performance of Cotton Stalk-Based Biochar
by Qiushuang Cui, Yong Huang, Xufei Ma, Sining Li, Ruyun Bai, Huan Li, Wen Liu and Hanyu Wei
Molecules 2024, 29(24), 5841; https://doi.org/10.3390/molecules29245841 - 11 Dec 2024
Viewed by 550
Abstract
In this research, we produced two types of biochar (BC) using cotton stalks as raw material and KOH as an activator, and compared their performance and adsorption mechanisms in the removal of tetracycline (TC) and methylene blue (MB) from wastewater. The results showed [...] Read more.
In this research, we produced two types of biochar (BC) using cotton stalks as raw material and KOH as an activator, and compared their performance and adsorption mechanisms in the removal of tetracycline (TC) and methylene blue (MB) from wastewater. The results showed that the biochar generated using both procedures formed pores that connected to the interior of the biochar and had extensive microporous and mesoporous structures. The molten salt approach produces biochar with a higher specific surface area, larger pore size, and higher pore volume than the impregnation method, with a maximum specific surface area of 3095 m2/g. KBCM-900 (the BC produced using the molten salt method at 900 °C) had a better adsorption effect on TC, with a clearance rate of more than 95% in 180 min and a maximum adsorption amount of 912.212 mg/g. The adsorption rates of the two BCs for MB did not differ significantly at low concentrations, but as the concentration increased, KBCI-900 (the BC generated by the impregnation method at 900 °C) exhibited better adsorption, with a maximum adsorption of 723.726 mg/g. The pseudo-second-order kinetic model and the Langmuir isotherm model may accurately describe the TC and MB adsorption processes of KBCI-900 and KBCM-900. The KBCI/KBCM-900 adsorption process combines physical and chemical adsorption, with the primary mechanisms being pore filling, π–π interactions, hydrogen bonding, and electrostatic interactions. As a result, biochar generated using the molten salt method is suitable for the removal of large-molecule pollutants such as TC, whereas biochar prepared using the impregnation method is suitable for the removal of small-molecule dyes such as MB. Full article
(This article belongs to the Special Issue Advanced Chemical Approaches and Technologies in Water Treatment)
Show Figures

Figure 1

Figure 1
<p>SEM of biochars: (<b>a</b>) KBCI-900; (<b>b</b>) KBCM-900.</p>
Full article ">Figure 2
<p>Nitrogen adsorption–desorption isotherms and pore size distribution of: (<b>a</b>) KBCI-900, (<b>b</b>) KBCM-900.</p>
Full article ">Figure 3
<p>FTIR of KBCI/KBCM-900 before and after adsorption: (<b>a</b>) KBCI-900; (<b>b</b>) KBCM-900; (<b>c</b>) Raman plots of KBCI-900 and KBCM-900.</p>
Full article ">Figure 4
<p>XPS images of BC: (<b>a</b>,<b>b</b>) KBCI-900 before and after adsorption; (<b>c</b>,<b>d</b>) KBCM-900 before and after adsorption.</p>
Full article ">Figure 5
<p>Adsorption performance of KBCI-900 and KBCM-900 on TC and MB.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Adsorption properties at different pH values; (<b>c</b>,<b>d</b>) zeta potentials.</p>
Full article ">Figure 7
<p>Kinetic modeling of TC and MB adsorption of biochar: (<b>a</b>) KBCI-900; (<b>b</b>) KBCM-900.</p>
Full article ">Figure 8
<p>BC to TC and MB internal diffusion model plots: (<b>a</b>) KBCI-900; (<b>b</b>) KBCM-900.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>c</b>) Adsorption isotherms of TC and MB on KBCI-900 and KBCM-900; (<b>b</b>,<b>d</b>) plots of separation factor (RL) versus initial TC concentration.</p>
Full article ">
Back to TopTop