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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline

Article Types

Countries / Regions

Search Results (202)

Search Parameters:
Keywords = Ob-R

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 20891 KiB  
Article
Efficient Photocatalytic Reduction of Hexavalent Chromium by NiCo2S4/BiOBr Heterogeneous Photocatalysts
by Shumeng Qin, Ruofan Xu, Qiu Jin, Sen Wang, Yi Ren, Yulin Huang, Ziye Zheng, Lihui Xiao, Dong Zhai, Shuguang Wang and Zuoli He
Coatings 2024, 14(12), 1492; https://doi.org/10.3390/coatings14121492 - 27 Nov 2024
Viewed by 500
Abstract
For typical Cr(VI)-containing industrial wastewater, more efficient water treatment technologies need to be used to ensure that Cr(VI) concentrations are reduced to safe levels before discharge. Photocatalytic technology is highly efficient, environmentally friendly, and has been extensively used to address this demand. Herein, [...] Read more.
For typical Cr(VI)-containing industrial wastewater, more efficient water treatment technologies need to be used to ensure that Cr(VI) concentrations are reduced to safe levels before discharge. Photocatalytic technology is highly efficient, environmentally friendly, and has been extensively used to address this demand. Herein, heterogeneous NiCo2S4/BiOBr photocatalysts with different ratios were prepared using a solvothermal method. When compared with pure NiCo2S4 and BiOBr, the NiCo2S4/BiOBr-30 had significantly increased adsorption capacity and visible-light-driven photocatalytic reduction activity for Cr(VI) removal. The improved adsorption performance of the NiCo2S4/BiOBr-30 was mainly due to its increased specific surface area, and the enhanced photocatalytic performance of the NiCo2S4/BiOBr-30 could be attributed to the improved separation and transfer of photogenerated carriers at the interface. Lastly, a possible enhanced photocatalytic Cr(VI) reduction mechanism of the NiCo2S4/BiOBr heterostructure was developed. Full article
(This article belongs to the Special Issue Advanced Materials and Coatings for Photocatalytic Applications)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of NiCo<sub>2</sub>S<sub>4</sub>, BiOBr, and NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30.</p>
Full article ">Figure 2
<p>SEM images, N<sub>2</sub> adsorption–desorption isotherms, and pore size distribution curves of (<b>a</b>) NiCo<sub>2</sub>S<sub>4</sub>, (<b>b</b>) BiOBr, and (<b>c</b>) NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30.</p>
Full article ">Figure 3
<p>The structures and density of states (DOS) plots of the (<b>a</b>) BiOBr, (<b>b</b>) NiCo<sub>2</sub>S<sub>4</sub>, and (<b>c</b>) BiOBr/NiCo<sub>2</sub>S<sub>4</sub>.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) HRTEM images and (<b>d</b>–<b>i</b>) elemental mapping images of NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30.</p>
Full article ">Figure 5
<p>(<b>a</b>) Full-scan XPS spectra. (<b>b</b>–<b>f</b>) High-resolution XPS spectra of (<b>b</b>) Ni 2p, (<b>c</b>) Co 2p, (<b>d</b>) Bi 4f and S 2p, (<b>e</b>) O 1s, and (<b>f</b>) Br 3d for NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30.</p>
Full article ">Figure 6
<p>(<b>a</b>) Test procedure for the photocatalytic reduction of Cr(VI). (<b>b</b>) Time profiles and (<b>c</b>) removal rate of the photocatalytic Cr(VI) reduction over NiCo<sub>2</sub>S<sub>4</sub>/BiOBr composite samples. (<b>d</b>–<b>f</b>) Comparisons of the Cr(VI) removal performances: (<b>d</b>) time profiles of the Cr(VI) removal, (<b>e</b>) time profiles of the photocatalytic Cr(VI) reduction, and (<b>f</b>) fitted curves for the kinetics of the first-stage reaction.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>f</b>) Cr(VI) removal performances of NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30 under visible light with (<b>a</b>,<b>d</b>) different pHs, (<b>b</b>,<b>e</b>) different anions (10 mM), and (<b>c</b>,<b>f</b>) different cations (10 mM). (<b>g</b>,<b>h</b>) Radical quenching experiment and corresponding Cr(VI) removal rates of NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30 composite. (<b>i</b>) Cyclic experiment with NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30.</p>
Full article ">Figure 8
<p>(<b>a</b>) Transient photocurrent response, (<b>b</b>) Nyquist plot and equivalent analog circuit diagram of the photocatalysts.</p>
Full article ">Figure 9
<p>(<b>a</b>) UV–Vis DRS spectra of the NiCo<sub>2</sub>S<sub>4</sub>, BiOBr, and NiCo<sub>2</sub>S<sub>4</sub>/BiOBr-30. (<b>b</b>) Kubelka–Munk plots for determining the bandgap energies. Mott–Schottky spectra of the (<b>c</b>) NiCo<sub>2</sub>S<sub>4</sub> and (<b>d</b>) BiOBr.</p>
Full article ">Figure 10
<p>Schematic diagram of the photocatalytic reduction of Cr(VI) by the NiCo<sub>2</sub>S<sub>4</sub>/BiOBr heterogeneous photocatalysts.</p>
Full article ">
14 pages, 2934 KiB  
Article
Detecting Honey Adulteration: Advanced Approach Using UF-GC Coupled with Machine Learning
by Irene Punta-Sánchez, Tomasz Dymerski, José Luis P. Calle, Ana Ruiz-Rodríguez, Marta Ferreiro-González and Miguel Palma
Sensors 2024, 24(23), 7481; https://doi.org/10.3390/s24237481 - 23 Nov 2024
Viewed by 506
Abstract
This article introduces a novel approach to detecting honey adulteration by combining ultra-fast gas chromatography (UF-GC) with advanced machine learning techniques. Machine learning models, particularly support vector regression (SVR) and least absolute shrinkage and selection operator (LASSO), were applied to predict adulteration in [...] Read more.
This article introduces a novel approach to detecting honey adulteration by combining ultra-fast gas chromatography (UF-GC) with advanced machine learning techniques. Machine learning models, particularly support vector regression (SVR) and least absolute shrinkage and selection operator (LASSO), were applied to predict adulteration in orange blossom (OB) and sunflower (SF) honeys. The SVR model achieved R2 values above 0.90 for combined honey types. Treating OB and SF honeys separately resulted in a significant accuracy improvement, with R2 values exceeding 0.99. LASSO proved especially effective when honey types were treated individually. The integration of UF-GC with machine learning not only provides a reliable method for detecting honey adulteration, but also sets a precedent for future research in the application of this technique to other food products, potentially enhancing food authenticity across the industry. Full article
Show Figures

Figure 1

Figure 1
<p>Circular dendrogram resulting from the hierarchical cluster analysis (HCA) of the dataset (D<sub>44<span class="html-italic">x</span>20002</sub>) with UF-GC; The names of the honey samples are colored according to their botanical origin: sunflower (purple) and orange blossom (orange). The four main clusters have been colored and labeled with letters A, B, C, and D. The average method with Euclidean distances was used.</p>
Full article ">Figure 2
<p>Score plot of PCA for orange blossom and sunflower honey.</p>
Full article ">
18 pages, 3764 KiB  
Article
Multifractal Analysis of Standardized Precipitation Evapotranspiration Index in Serbia in the Context of Climate Change
by Tatijana Stosic, Ivana Tošić, Irida Lazić, Milica Tošić, Lazar Filipović, Vladimir Djurdjević and Borko Stosic
Sustainability 2024, 16(22), 9857; https://doi.org/10.3390/su16229857 - 12 Nov 2024
Viewed by 677
Abstract
A better understanding of climate change impact on dry/wet conditions is crucial for agricultural planning and the use of renewable energy, in terms of sustainable development and preservation of natural resources for future generations. The objective of this study was to investigate the [...] Read more.
A better understanding of climate change impact on dry/wet conditions is crucial for agricultural planning and the use of renewable energy, in terms of sustainable development and preservation of natural resources for future generations. The objective of this study was to investigate the impact of climate change on temporal fluctuations of dry/wet conditions in Serbia on multiple temporal scales through multifractal analysis of the standardized precipitation evapotranspiration index (SPEI). We used the well-known method of multifractal detrended fluctuation analysis (MFDFA), which is suitable for the analysis of scaling properties of nonstationary temporal series. The complexity of the underlying stochastic process was evaluated through the parameters of the multifractal spectrum: position of maximum α0 (persistence), spectrum width W (degree of multifractality) and skew parameter r dominance of large/small fluctuations). MFDFA was applied on SPEI time series for the accumulation time scale of 1, 3, 6 and 12 months that were calculated using the high-resolution meteorological gridded dataset E-OBS for the period from 1961 to 2020. The impact of climate change was investigated by comparing two standard climatic periods (1961–1990 and 1991–2020). We found that all the SPEI series show multifractal properties with the dominant contribution of small fluctuations. The short and medium dry/wet conditions described by SPEI-1, SPEI-3, and SPEI-6 are persistent (0.5<α0<1); stronger persistence is found at higher accumulation time scales, while the SPEI-12 time series is antipersistent (0<α01<0.5). The degree of multifractality increases from SPEI-1 to SPEI-6 and decreases for SPEI-12. In the second period, the SPEI-1, SPEI-3, and SPEI-6 series become more persistent with weaker multifractality, indicating that short and medium dry/wet conditions (which are related to soil moisture and crop stress) become easier to predict, while SPEI-12 changed toward a more random regime and stronger multifractality in the eastern and central parts of the country, indicating that long-term dry/wet conditions (related to streamflow, reservoir levels, and groundwater levels) become more difficult for modeling and prediction. These results indicate that the complexity of dry/wet conditions, in this case described by the multifractal properties of the SPEI temporal series, is affected by climate change. Full article
(This article belongs to the Special Issue The Future of Water, Energy and Carbon Cycle in a Changing Climate)
Show Figures

Figure 1

Figure 1
<p>Position of Serbia in Europe and map of Serbia with its orography and major rivers.</p>
Full article ">Figure 2
<p>Multifractal spectra for SPEI-1 for a sample grid point at latitude 43.15 and longitude 22.45, corresponding to the city of Pirot for the two accumulation periods 1961–1990 and 1991–2020. In the top row, the fluctuation function versus scale on the log-log plot is displayed, together with linear fits for different <math display="inline"><semantics> <mrow> <mi>q</mi> </mrow> </semantics></math> values. In the middle row, the functions <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>(</mo> <mi>q</mi> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mfenced separators="|"> <mrow> <mi>q</mi> </mrow> </mfenced> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mfenced separators="|"> <mrow> <mi>α</mi> </mrow> </mfenced> </mrow> </semantics></math> are shown (lines serve to guide the eye), and in the bottom row, the maps of the parameter <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> are shown, emphasizing the chosen sample grid point position (bold squares in the southeast) for the period 1961–1990 (<b>bottom left</b>) and 1991–2020 (<b>bottom right</b>).</p>
Full article ">Figure 3
<p>Mutifractal parameter <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> for SPEI-1, SPEI-3, SPEI-6 and SPEI-12 across Serbia for the periods 1961–1990 and 1991–2020. To emphasize the difference between the parameters among the two periods, the range of the color bar is adjusted to cover (roughly) <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>2</mn> </mrow> </semantics></math> standard deviations of the mean of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> for each SPEI accumulation period.</p>
Full article ">Figure 4
<p>Mutifractal parameter <math display="inline"><semantics> <mrow> <mi>W</mi> </mrow> </semantics></math> for SPEI-1, SPEI-3, SPEI-6 and SPEI-12 across Serbia for the periods 1961–1990 and 1991–2020.</p>
Full article ">Figure 5
<p>Mutifractal parameter <math display="inline"><semantics> <mrow> <mi>r</mi> </mrow> </semantics></math> for SPEI-1, SPEI-3, SPEI-6 and SPEI-12 across Serbia for the periods 1961–1990 and 1991–2020.</p>
Full article ">
15 pages, 3198 KiB  
Article
Inhalable Anti-EGFR Antibody-Conjugated Osimertinib Liposomes for Non-Small Cell Lung Cancer
by Apoorva Daram, Shruti S. Sawant, Dhwani A. Mehta, Carlos A. Sanhueza and Nitesh K. Kunda
Pharmaceutics 2024, 16(11), 1444; https://doi.org/10.3390/pharmaceutics16111444 - 12 Nov 2024
Viewed by 1022
Abstract
Background: Non-small cell lung cancer (NSCLC) is a leading cause of cancer deaths globally. The most extensive treatment is Tyrosine Kinase Inhibitors (TKIs) that target epidermal growth factor receptor (EGFR) overexpression. Osimertinib, a third-generation TKI is approved to target EGFR exon 19 [...] Read more.
Background: Non-small cell lung cancer (NSCLC) is a leading cause of cancer deaths globally. The most extensive treatment is Tyrosine Kinase Inhibitors (TKIs) that target epidermal growth factor receptor (EGFR) overexpression. Osimertinib, a third-generation TKI is approved to target EGFR exon 19 deletions or exon 21 L858R mutations. However, resistance is inevitable due to emergence of triple mutations (sensitizing mutations, T790M and C797S). To overcome this challenge, a combinatorial approach was used wherein Osimertinib liposomes were conjugated with cetuximab (CTX), an anti-EGFR monoclonal antibody, to improve drug efficacy and delivery. Additionally, pulmonary administration was employed to minimize systemic toxicity and achieve high lung concentrations. Methods: Osimertinib liposomes (OB-LPs) were prepared using thin film hydration method and immunoliposomes (CTX-OB-LPs) were prepared by conjugating the OB-LPs surface with CTX. Liposomes were characterized for particle size, zeta-potential, drug loading, antibody conjugation efficiency, in vitro drug release, and aerosolization performance. Further, the in vitro efficacy of immunoliposomes was evaluated in H1975 cell line. Results: Immunoliposomes exhibited a particle size of 150 nm, high antibody conjugation efficiency (87%), efficient drug release, and excellent aerosolization properties with an aerodynamic diameter of 3 μm and fine particle fraction of 88%. Furthermore, in vitro studies in H1975 cells showed enhanced cytotoxicity with CTX-OB-LPs displaying 1.7-fold reduction and 1.2-fold reduction in IC50 compared to Osimertinib and OB-LPs, respectively. The CTX-OB-LPs also significantly reduced tumor cell migration and colonization compared to Osimertinib and OB-LPs. Conclusions: These successful results for EGFR-targeting inhalable immunoliposomes exhibited potential for contributing to greater anti-tumor efficacy for the treatment of non-small cell lung cancer. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for liposomal formulation preparation.</p>
Full article ">Figure 2
<p>The XRD diffractogram of OB, blank formulation (blank LPs), unconjugated OB liposomes (OB-LPs), and conjugated OB liposomes (CTX-OB-LPs). The crystalline peaks of OB are absent in the XRD diffractogram of OB-LPs and CTX-OB-LPs, indicating drug encapsulation within the liposomes.</p>
Full article ">Figure 3
<p>(<b>a</b>) Cumulative release profile for OB from immunoliposomes (CTX-OB-LPs) in phosphate buffer saline (PBS), pH 7.4. Data represents mean ± SD (<span class="html-italic">n</span> = 4). (<b>b</b>) In vitro aerosol deposition profile represented as percentage of drug deposited on each stage of next generation impactor (NGI). Data represents mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Kinetic analysis using SPR for the (<b>a</b>) binding of antibodies (CTX) to the EGFR protein and (<b>b</b>) binding of CTX-OB-LPs to the EGFR protein. Data were fitted using the TraceDrawer software at various concentrations injected at 20 μL/min over EGFR immobilized on the sensor’s surface. Data represents mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Cytotoxicity studies after 72 h treatment, as determined using the MTT assay in the H1975 cell line. (<b>a</b>) Blank liposomes (blank LPs) and CTX-conjugated liposomes (CTX-LPs); (<b>b</b>) OB, unconjugated OB liposomes (OB-LPs), and conjugated OB liposomes (CTX-OB-LPs). Data represents mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>Colony-forming ability of H1975 cells under treatment for 72 h, followed by a 10-day incubation in fresh media. (<b>a</b>) Quantitative analysis of the clonogenic nature of the H1975 cells after treatment with OB, unconjugated OB liposomes (OB-LPs), and conjugated OB liposomes (CTX-OB-LPs). (<b>b</b>) Images of the colonies after crystal violet staining. The data are expressed as % colony growth versus the respective treatment. Data represents mean ± SD (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.0001 and ** <span class="html-italic">p</span> &lt; 0.001; ns—non-significant.</p>
Full article ">Figure 7
<p>(<b>a</b>) Scratch assay analysis of the H1975 cell line, shown as % of wound healing over time after treatment with OB, unconjugated OB liposomes (OB-LPs), and conjugated OB liposomes (CTX-OB-LPs). Data represents mean ± SD (<span class="html-italic">n</span> = 3). **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01, and * <span class="html-italic">p</span> &lt; 0.05; ns—non-significant. (<b>b</b>) Effect of OB and liposomal formulations (OB-LPs and CTX-OB-LPs) on the metastatic potential of the H1975 cell line. Representative microscopic images of the scratch after the following treatment times are provided: 0 h, 12 h, 24 h, and 48 h. Scale bar 400 µm.</p>
Full article ">Figure 8
<p>Stability data for CTX-OB-LPs when stored at 4 °C. (<b>a</b>) % entrapment efficiency (EE); (<b>b</b>) drug content; (<b>c</b>) particle size; (<b>d</b>) zeta potential.</p>
Full article ">
14 pages, 422 KiB  
Article
Continuous Glucose Monitor Metrics That Predict Neonatal Adiposity in Early and Later Pregnancy Are Higher in Obesity Despite Macronutrient-Controlled Eucaloric Diets
by Teri L. Hernandez, Sarah S. Farabi, Rachael E. Van Pelt, Nicole Hirsch, Emily Z. Dunn, Elizabeth A. Haugen, Melanie S. Reece, Jacob E. Friedman and Linda A. Barbour
Nutrients 2024, 16(20), 3489; https://doi.org/10.3390/nu16203489 - 15 Oct 2024
Viewed by 810
Abstract
Background: Fasting glucose is higher in pregnancies with obesity (OB); less is known about postprandial (PP) and nocturnal patterns when the diet is eucaloric and fixed or about the continuous-glucose-monitor (CGM) metrics that predict neonatal adiposity (NB%fat). We hypothesized that continuous glucose monitors [...] Read more.
Background: Fasting glucose is higher in pregnancies with obesity (OB); less is known about postprandial (PP) and nocturnal patterns when the diet is eucaloric and fixed or about the continuous-glucose-monitor (CGM) metrics that predict neonatal adiposity (NB%fat). We hypothesized that continuous glucose monitors (CGMs) would reveal higher glycemia in OB vs. normal weight (NW) during Early (14–16 weeks) and Later (26–28 weeks) gestation despite macronutrient-controlled eucaloric diets and elucidate unique predictors of NB%fat. Methods: In a prospective, parallel-group comparative study, a eucaloric diet (NW: 25 kcal/kg; OB: 30 kcal/kg) was provided (50% carbohydrate [20% simple/30% complex; of total calories], 35% fat, 15% protein) to Early and Later gestation groups wearing a blinded CGM for three days. CGM metrics (mean fasting; 1 h and 2 h PP; daytime and nocturnal glucose; percent time-in-range (%TIR: 63–140 mg/dL); PP excursions; and area-under-the-curve [AUC]) were interrogated between groups and as predictors of NB%fat by dual X-ray absorptiometry(DXA). Results: Fifty-four women with NW (BMI: 23 kg/m2; n = 27) and OB (BMI: 32; n = 27) provided their informed consent to participate. Early, the daytime glucose was higher in OB vs. NW (mean ± SEM) (91 ± 2 vs. 85 ± 2 mg/dL, p = 0.017), driven by 2 h PP glucose (95 ± 2 vs. 88 ± 2, p = 0.004). Later, those with OB exhibited higher nocturnal (89 ± 2 vs. 81 ± 2), daytime (95 ± 2 vs. 87 ± 2), 1 h (109 ± 3 vs. 98 ± 2), and 2 h PP (101 ± 3 vs. 92 ± 2) glucose (all p < 0.05) but no difference in %TIR (95–99%). Postprandial peak excursions for all meals were markedly blunted in both the Early (9–19 mg/dL) and Later (15–26 mg/dL). In OB, the Later group’s 24 h AUC was correlated with NB%fat (r = 0.534, p = 0.02). Despite similar weight gain, infants of OB had higher birthweight (3528 ± 107 vs. 3258 ± 74 g, p = 0.037); differences in NB%fat did not reach statistical significance (11.0 vs. 8.9%; p > 0.05). Conclusions: Despite macronutrient-controlled eucaloric diets, pregnancies with OB had higher glycemia Early and Later in gestation; the Later 24 h glucose AUC correlated with NB%fat. However, glycemic patterns were strikingly lower than current management targets. Full article
(This article belongs to the Special Issue Featured Articles on Nutrition and Obesity Management (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Patterns of 24 h glycemia measured by CGM in participants with NW and OB, both <span class="html-italic">Early</span> (14–16 weeks, Panel (<b>A</b>)) and <span class="html-italic">Later</span> (26–28 weeks, Panel (<b>B</b>)) in pregnancy. Gray and black dashed lines show mean nocturnal and daytime glucose between the groups.</p>
Full article ">
13 pages, 8871 KiB  
Article
Thin Films of Bismuth Oxyhalides (BiOX, X = Cl, Br, I) Deposited by Thermal Evaporation for the Decontamination of Water and Air by Photocatalysis
by Enrique López-Cuéllar, Azael Martínez-de la Cruz, Rodolfo Morales-Ibarra, Marco Garza-Navarro and José Olivares-Cortez
Catalysts 2024, 14(10), 716; https://doi.org/10.3390/catal14100716 - 14 Oct 2024
Viewed by 678
Abstract
Thin films of BiOCl, BiOBr, and BiOI (BiOX) were deposited by thermal evaporation for their potential application in the decontamination of water and air through their photocatalytic activity, which was compared among the three. The BiOX thin films were subjected to characterization through [...] Read more.
Thin films of BiOCl, BiOBr, and BiOI (BiOX) were deposited by thermal evaporation for their potential application in the decontamination of water and air through their photocatalytic activity, which was compared among the three. The BiOX thin films were subjected to characterization through X-ray diffraction, high-resolution transmission electron microscopy, and scanning electron microscopy. Additionally, the optical properties were determined from the diffuse reflectance spectrum obtained with a spectrophotometer. To assess the efficacy of the semiconductor films in water decontamination, the evolution of rhodamine B discoloration and its mineralization was monitored by measuring total organic carbon. The decontaminating activity in the air was evaluated in a gas reactor, measuring the conversion of NOx-type gases. The results demonstrated that the thin films of the three oxides exhibited decontaminating photocatalytic activity in both water and air. However, notable distinctions were observed in the photocatalytic activities of the three bismuth oxyhalides in water, while in air, they exhibited similarities. In aqueous environments, the mineralization percentages exhibited notable variation after 96 h, with the BiOBr film displaying a value of 9.2%/mg and the BiOCl film a value of 3.9%/mg. In contrast, the NO conversion rate in the air was approximately 0.6%/mg for the three oxyhalide films. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>Thin films of bismuth oxyhalides.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) Diffractograms of the BiOCl, BiOBr, and BiOI films and of the powders used for the deposition of the films, respectively.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>–<b>c</b>) Diffractograms of the BiOCl, BiOBr, and BiOI films and of the powders used for the deposition of the films, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) TEM images of the BiOCl, BiOBr, and BiOI films and (<b>d</b>–<b>f</b>) HRTEM images and its FFT of the films, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) SEM micrographs in perpendicular view of thin films of (<b>a</b>) BiOCl, (<b>b</b>) BiOBr, and (<b>c</b>) BiOI. (<b>d</b>–<b>f</b>) Lateral view of thin films of (<b>d</b>) BiOCl, (<b>e</b>) BiOBr, and (<b>f</b>) BiOI.</p>
Full article ">Figure 5
<p>Absorption spectra of BiOCl, BiOBr, and BiOI in film form.</p>
Full article ">Figure 6
<p>Tauc curve for Eg of (<b>a</b>) BiOCl, (<b>b</b>) BiOBr, and (<b>c</b>) BiOI films.</p>
Full article ">Figure 7
<p>Bleaching activity versus time for BiOCl, BiOBr, and BiOI films.</p>
Full article ">Figure 8
<p>Mineralization percentages achieved by bismuth oxyhalide films.</p>
Full article ">Figure 9
<p>NO conversion rates achieved by each oxide film.</p>
Full article ">
15 pages, 5659 KiB  
Article
In Situ Metallic Bi-Modified (110)BiOBr Nanosheets with Surface Plasmon Resonance Effect for Enhancing Photocatalytic Performance Despite of Larger Optical Band Gap
by Yunhe Mu, Hongxue Chu, Hougang Fan, Xin Li, Xiaoyan Liu, Lili Yang, Maobin Wei and Huilian Liu
Catalysts 2024, 14(9), 654; https://doi.org/10.3390/catal14090654 - 23 Sep 2024
Viewed by 595
Abstract
BiOBr with different preferred growth orientation facets would show a different photocatalytic performance. When decorated in situ with metallic Bi nanoparticles, Bi/BiOBr would commonly display an enhanced photocatalytic performance. In this paper, the BiOBr nanoplates with preferred growth orientation (102) facet and (110) [...] Read more.
BiOBr with different preferred growth orientation facets would show a different photocatalytic performance. When decorated in situ with metallic Bi nanoparticles, Bi/BiOBr would commonly display an enhanced photocatalytic performance. In this paper, the BiOBr nanoplates with preferred growth orientation (102) facet and (110) facet were first synthesized using a hydrothermal method. Then, some metallic Bi nanoparticles were modified in situ onto the (110)BiOBr nanoplates, which was expected to show a much more enhanced photocatalytic performance. All samples were characterized using XRD, FE-SEM, TEM, N2 adsorption–desorption, UV–vis and XPS. FE-SEM and TEM images showed that the grain size of the metallic Bi particles was about 5 nm to 10 nm. UV–vis spectra showed that, after some metallic Bi nanoparticles were modified on (110)BiOBr nanoplates, the light absorbance in the visible light region at 400–700 nm became stronger and their optical band gap became larger. N2 adsorption–desorption tests showed that the Bi(x)/(110)BiOBr nanosheets possessed larger specific surface areas than that of the (102)BiOBr and (110)BiOBr nanoplates. The XPS results showed that Bi(x)/(110)BiOBr contained more oxygen vacancies and a more negative value of the conduction band minimum. The photocatalytic performance of (102)BiOBr, (110)BiOBr and Bi(x)/(110)BiOBr were tested in the photocatalytic degradation of rhodamine B under visible light irradiation for 2 h; their photocatalytic efficiency was 45%, 75% and 80%, respectively. In comparison to (102)BiOBr, (110)BiOBr exhibited much higher photocatalytic activity, while for Bi(x)/(110)BiOBr, despite the surface Plasmon resonance effect, a larger specific surface area and more oxygen vacancies, the enhancement of the efficiency was limited, which might have resulted from the larger optical band gap. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>XRD pattern of (<b>A</b>) the (102)BiOBr and (110)BiOBr samples, (<b>B</b>) Bix/(110)BiOBr samples.</p>
Full article ">Figure 2
<p>SEM images of (<b>A</b>) (102)BiOBr, (<b>B</b>) (110)BiOBr, (<b>C</b>) Bi(0.5 mmol)/(110)BiOBr, (<b>D</b>) Bi(1.0 mmol)/(110)BiOBr, (<b>E</b>) Bi(1.5 mmol)/(110)BiOBr and TEM images of (<b>F</b>), (<b>G</b>) Bi(1.0 mmol)/(110)BiOBr.</p>
Full article ">Figure 2 Cont.
<p>SEM images of (<b>A</b>) (102)BiOBr, (<b>B</b>) (110)BiOBr, (<b>C</b>) Bi(0.5 mmol)/(110)BiOBr, (<b>D</b>) Bi(1.0 mmol)/(110)BiOBr, (<b>E</b>) Bi(1.5 mmol)/(110)BiOBr and TEM images of (<b>F</b>), (<b>G</b>) Bi(1.0 mmol)/(110)BiOBr.</p>
Full article ">Figure 2 Cont.
<p>SEM images of (<b>A</b>) (102)BiOBr, (<b>B</b>) (110)BiOBr, (<b>C</b>) Bi(0.5 mmol)/(110)BiOBr, (<b>D</b>) Bi(1.0 mmol)/(110)BiOBr, (<b>E</b>) Bi(1.5 mmol)/(110)BiOBr and TEM images of (<b>F</b>), (<b>G</b>) Bi(1.0 mmol)/(110)BiOBr.</p>
Full article ">Figure 3
<p>The photocatalytic performance of removing RhB using (<b>A</b>) the (102)BiOBr and (110)BiOBr samples, (<b>B</b>) Bix/(110)BiOBr samples.</p>
Full article ">Figure 4
<p>N<sub>2</sub> adsorption–desorption curves of (<b>A</b>) (102)BiOBr and (110)BiOBr; (<b>B</b>) Bi(0.5 mmol)/ (110)BiOBr, Bi(1.0 mmol)/(110)BiOBr and Bi(1.5 mmol)/(110)BiOBr.</p>
Full article ">Figure 5
<p>UV-vis absorption spectra of all the samples.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>C</b>) XPS spectra of the selected three samples and (<b>D</b>–<b>F</b>) the deconvolution results of the selected three samples in high resolution O 1s XPS spectra.</p>
Full article ">Figure 7
<p>(<b>A</b>) The photocurrents and (<b>B</b>) PL spectra of all the samples.</p>
Full article ">Figure 8
<p>Photocatalytic removal of RhB with four scavengers by (<b>A</b>) (102)BiOBr, (<b>B</b>) (110)BiOBr and (<b>C</b>) Bi(1.0mmol)/(110)BiOBr.</p>
Full article ">Figure 9
<p>The schematic diagram of the role of the SPR effect between metallic Bi and BiOBr.</p>
Full article ">Figure 10
<p>The schematic diagram of energy level and the photogenerated electrons transfer of the selected three samples.</p>
Full article ">
16 pages, 1161 KiB  
Article
Characterizing Human Peripheral Blood Lymphocyte Phenotypes and Their Correlations with Body Composition in Normal-Weight, Overweight, and Obese Healthy Young Adults
by Irina-Bianca Kosovski, Cristina Nicoleta Ciurea, Dana Ghiga, Naomi-Adina Ciurea, Adina Huțanu, Florina Ioana Gliga and Anca Bacârea
Medicina 2024, 60(9), 1523; https://doi.org/10.3390/medicina60091523 - 18 Sep 2024
Viewed by 791
Abstract
Background and Objectives: Obesity-associated chronic low-grade inflammation supports various systemic alterations. In this descriptive study, 122 apparently healthy adults aged 20 to 35 years were voluntarily included and classified based on body mass index (BMI) as normal-weight (NW), overweight (OW), and obese (OB). [...] Read more.
Background and Objectives: Obesity-associated chronic low-grade inflammation supports various systemic alterations. In this descriptive study, 122 apparently healthy adults aged 20 to 35 years were voluntarily included and classified based on body mass index (BMI) as normal-weight (NW), overweight (OW), and obese (OB). This study aims to characterize peripheral blood (PB) lymphocyte (Ly) phenotypes and investigate their correlations with body composition indices (BCIs) in healthy young adults. Materials and Methods: The following BCIs were measured: waist circumference, hip circumference, height, waist-to-hip ratio, waist-to-height ratio, total body fat mass, visceral fat level, weight, and BMI. White blood cell count (WBC), Ly absolute count, serum TNF-α, and IFN-γ were quantified. Ly subpopulations were analyzed as follows: total TLy (TTLy—CD45+CD3+), early activated TLy (EATLy—CD45+3+69+), total NKLy (TNKLy—CD45+CD3CD56+CD16+), NKdim (low expression of CD56+), NKbright (high expression of CD56+), BLy (CD45+CD3CD19+), T helper Ly (ThLy—CD45+CD3+CD4+), and T cytotoxic Ly (TcLy—CD45+CD3+CD8+). Results: Higher BMI has significantly higher WBC and BLy (p < 0.0001; p = 0.0085). EATLy significantly decreased from NW to OB (3.10—NW, 1.10—OW, 0.85—OB, p < 0.0001). Only EATLy exhibited significant negative correlations with all the BCIs. A significantly higher TNF-α was observed in the OW and OB groups compared to the NW group. IFN-γ increased linearly but nonsignificantly with BMI. TTLy showed a nonsignificant positive correlation with both IFN-γ and TNF-α, while EATLy showed a negative correlation, significant only for IFN-γ. NKLy subpopulations exhibited a consistent negative correlation with TNF-α, significant only for NKdim (p = 0.0423), and a nonsignificant consistent positive correlation with IFN-γ. A nonsignificant negative correlation between age and both TNKLy (r = −0.0927) and NKdim (r = −0.0893) cells was found, while a positive correlation was found with NKbright (r = 0.0583). Conclusions: In conclusion, the baseline immunological profile of PB is influenced by excessive adipose tissue in healthy young adults. Full article
(This article belongs to the Section Epidemiology & Public Health)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Correlation between total T lymphocytes (<b>A</b>) and early activated T lymphocyte populations (<b>B</b>) with IFN-γ and TNF-α. IFN-γ—interferon-γ; TNF-α—tumor necrosis factor-α; 95%CI—95% confidence interval; PBMCs—peripheral blood mononuclear cells. All correlations were analyzed using the Pearson test.</p>
Full article ">Figure 2
<p>Correlation between NKLy subsets (<b>A</b>)—Total NK lymphocytes; (<b>B</b>)—NK<sup>dim</sup> lymphocytes; (<b>C</b>)—NK<sup>bright</sup> lymphocytes) and IFN-γ and TNF-α. IFN-γ—interferon-γ; TNF-α—tumor necrosis factor-α; 95%CI—95% confidence interval; PBMCs—peripheral blood mononuclear cells. All correlations were analyzed using the Pearson test.</p>
Full article ">Figure 3
<p>Correlation between NKLy populations and age. IFN-γ—interferon-γ; TNF-α—tumor necrosis factor-α; 95%CI—95% confidence interval; PBMCs—peripheral blood mononuclear cells. All correlations were analyzed using the Pearson test.</p>
Full article ">
12 pages, 645 KiB  
Article
Degree of Food Processing (NOVA Classification) and Blood Pressure in Women with Overweight and Obesity
by Amanda F. de Sousa, Jéssica de O. Campos, Débora K. da S. Oliveira, Jéssica G. Pereira, Márcia J. do E. Santo, Viviane de O. N. Souza, Aiany C. Simões-Alves and João H. Costa-Silva
Obesities 2024, 4(3), 353-364; https://doi.org/10.3390/obesities4030028 - 4 Sep 2024
Viewed by 803
Abstract
(1) Background: We aimed to associate the degree of food processing with blood pressure levels in adult women. (2) Methods: A cross-sectional study was carried out on 85 adult women. The participants were subdivided into three groups: normal weight (NW: 27.05%), overweight (OW: [...] Read more.
(1) Background: We aimed to associate the degree of food processing with blood pressure levels in adult women. (2) Methods: A cross-sectional study was carried out on 85 adult women. The participants were subdivided into three groups: normal weight (NW: 27.05%), overweight (OW: 34.1%) and obesity (OB: 38.8%). Their anthropometric parameters, food consumption and blood pressure (BP) were evaluated. The groups were compared using one-way ANOVA or the Kruskal–Wallis test, and correlations were established using Spearman’s correlation, partial correlations (adjusted for age, medications and pathologies) and simple linear regression. Significance was set at p < 0.05. (3) Results: Women with obesity had higher systolic and diastolic blood pressure (SBP = NW: 106.5 ± 11.6; OW: 111.60 ± 11.8; OB: 123.63 ± 14.0; p < 0.001 and DBP = NW: 66.5 ± 9.9; OW: 70.2 ± 8.7; OB: 80.6 ± 11.0; p < 0.001) and a lower consumption of unprocessed or minimally processed food (MPF) (NW: 0.25 ± 0.1; OW: 0.27 ± 0.09; OB: 0.21 ± 0.07; p = 0.027). Moreover, we found an inverse association among the consumption of MPF and diastolic blood pressure that remained after adjustments for covariates (r: −0.27; p = 0.01), suggesting that lower consumption of MPF is related to higher levels of DBP. (4) Conclusions: Our data suggest that women with obesity seem to have a lower consumption of MPF. In addition, MPF is negatively correlated with BP, suggesting an association with obesity and cardiovascular health. Full article
Show Figures

Figure 1

Figure 1
<p>Food consumption according to the degree of food processing in adult women classified according to BMI in Vitória de Santo Antão—PE, 2022. MPF: unprocessed/minimally processed food; PF: processed food; UPF: ultra-processed food; white bar: women with a normal weight; gray bar: overweight women; black bar: women with obesity. (#): When <span class="html-italic">p</span> &lt; 0.05 compared to the overweight group; values were expressed as means ± standard deviation of the mean. Kruskal–Wallis test with Dunn’s post hoc test.</p>
Full article ">Figure 2
<p>Correlation between body mass index and systolic and diastolic blood pressure in adult women in Vitória de Santo Antão—PE, 2022. BMI: body mass index; SBP: systolic blood pressure; DBP: diastolic blood pressure. r: Spearman’s correlation coefficient; <span class="html-italic">p</span>: value. Spearman’s correlation.</p>
Full article ">
14 pages, 4486 KiB  
Article
Efficient and Robust Photodegradation of Dichlorvos Pesticide by BiOBr/WO2.72 Nanocomposites with Type-I Heterojunction under Visible Light Irradiation
by Aoyun Meng, Wen Li, Zhen Li and Jinfeng Zhang
Catalysts 2024, 14(8), 548; https://doi.org/10.3390/catal14080548 - 21 Aug 2024
Viewed by 908
Abstract
In this study, we developed novel BiOBr/WO2.72 nanocomposites (abbreviated as BO/WO) and systematically investigated their photocatalytic degradation performance against the pesticide dichlorvos under visible light irradiation. The experimental results demonstrated that the BO/WO nanocomposites achieved an 85.4% degradation of dichlorvos within 80 [...] Read more.
In this study, we developed novel BiOBr/WO2.72 nanocomposites (abbreviated as BO/WO) and systematically investigated their photocatalytic degradation performance against the pesticide dichlorvos under visible light irradiation. The experimental results demonstrated that the BO/WO nanocomposites achieved an 85.4% degradation of dichlorvos within 80 min. In comparison, the BO alone achieved a degradation degree of 66.8%, and the WO achieved a degradation degree of 64.7%. Furthermore, the BO/WO nanocomposites retained 96% of their initial activity over five consecutive cycles, demonstrating exceptional stability. Advanced characterization techniques, such as high-resolution transmission electron microscopy (HRTEM), X-ray photoelectron spectroscopy (XPS), and electron paramagnetic resonance (EPR) confirmed the composition and catalytic mechanism of the composite material. The findings indicated that the BO/WO nanocomposites, through their optimized Type-I heterojunction structure, achieved efficient separation and transport of photogenerated electron–hole pairs, significantly enhancing the degree of degradation of organophosphate pesticides. This research not only propels the development of high-performance photocatalytic materials, but also provides innovative strategies and a robust scientific foundation for mitigating global organophosphate pesticide pollution, underscoring its substantial potential for environmental remediation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>a</b>) BO and, BO/WO nanocomposites, and WO; (<b>b</b>) FT-IR spectra of BO, BO/WO nanocomposites, and WO.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) BO, (<b>b</b>) WO, and (<b>c</b>) BO/WO; EDS spectra of (<b>d</b>) BO, (<b>e</b>) WO, and (<b>f</b>) BO/WO. TEM images of (<b>g</b>) BO, (<b>h</b>) WO, and (<b>i</b>) BO/WO (inset shows the HRTEM image of the BO/WO nanocomposites).</p>
Full article ">Figure 3
<p>XPS spectra: (<b>a</b>) Survey spectra; (<b>b</b>) Br 3d; (<b>c</b>) O 1s; (<b>d</b>) Bi 4f; (<b>e</b>) W 4f; (<b>f</b>) BET surface area of catalysts.</p>
Full article ">Figure 4
<p>(<b>a</b>) Photocatalytic dichlorvos degradation profiles of BO, (<b>b</b>) BO/WO nanocomposites, and (<b>c</b>) WO under visible light irradiation. Bottom panels show the corresponding degradation extents over time.</p>
Full article ">Figure 5
<p>Photocatalytic stability of (<b>a</b>) BO and, (<b>b</b>) BO/WO nanocomposites, and of (<b>c</b>) WO for the degradation of dichlorvos under visible light irradiation across multiple cycles.</p>
Full article ">Figure 6
<p>Degradation degree of BO/WO with (<b>a</b>) IPA, (<b>b</b>) ascorbic acid, (<b>c</b>) TEOA, and (<b>d</b>) KBrO<sub>3</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) UV-vis absorption spectra; (<b>b</b>) Tauc plot for WO; (<b>c</b>) Tauc plot for BO; (<b>d</b>) EPR spectra of BO/WO; (<b>e</b>) electrostatic potential of WO; (<b>f</b>) electrostatic potential of BO.</p>
Full article ">Figure 8
<p>(<b>a</b>) Photocurrent responses of the BO, WO, and BO/WO nanocomposites; (<b>b</b>) EIS Nyquist plots of the BO, WO, and BO/WO nanocomposites.</p>
Full article ">Figure 9
<p>The photocatalysis mechanism of the BO/WO nanocomposites under visible light.</p>
Full article ">Scheme 1
<p>Schematic illustration of the synthesis of BO/WO nanocomposites.</p>
Full article ">
12 pages, 6701 KiB  
Article
Construction of a Wood Nanofiber–Bismuth Halide Photocatalyst and Catalytic Degradation Performance of Tetracycline from Aqueous Solutions
by Jiarong She, Cuihua Tian, Yan Qing and Yiqiang Wu
Molecules 2024, 29(14), 3253; https://doi.org/10.3390/molecules29143253 - 10 Jul 2024
Viewed by 962
Abstract
Nanostructured bismuth oxide bromide (BiOBr) has attracted considerable attention as a visible light catalyst. However, its photocatalytic degradation efficiency is limited by its low specific surface area. In this study, a solvothermal approach was employed to synthesize BiOBr, which was subsequently loaded onto [...] Read more.
Nanostructured bismuth oxide bromide (BiOBr) has attracted considerable attention as a visible light catalyst. However, its photocatalytic degradation efficiency is limited by its low specific surface area. In this study, a solvothermal approach was employed to synthesize BiOBr, which was subsequently loaded onto cellulose nanofibers (CNFs) to obtain a bismuth halide composite catalyst. The performance of this catalyst in the removal of refractory organic pollutants such as tetracycline (TC) from solutions under visible light excitation was examined. Our results indicate that BiOBr/CNF effectively removes TC from the solution under light conditions. At a catalyst dosage of 100 mg/L, the removal efficiency for TC (with an initial concentration of 100 mg/L) was 94.2%. This study elucidates the relationship between the microstructure of BiOBr/CNF composite catalysts and their improved photocatalytic activity, offering a new method for effectively removing pollutants from water. Full article
(This article belongs to the Special Issue Advanced Materials in Photoelectrochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction patterns and (<b>b</b>) Fourier transform infrared spectra of the prepared samples.</p>
Full article ">Figure 2
<p>Scanning electron microscopy images of (<b>a</b>) cellulose nanofibers (CNF), (<b>b</b>) BiOBr/CNF.</p>
Full article ">Figure 3
<p>Transmission electron microscopy images of (<b>a</b>) CNF and (<b>b</b>) BiOBr/CNF.</p>
Full article ">Figure 4
<p>UV–vis diffused reflectance spectra of samples.</p>
Full article ">Figure 5
<p>The I-T curves (<b>a</b>) and the EIS response (<b>b</b>) of sample.</p>
Full article ">Figure 6
<p>Adsorption and degradation performance of tetracycline in various samples under (<b>a</b>) light avoidance, (<b>b</b>) visible light exposure, and (<b>c</b>) UV light exposure.</p>
Full article ">Figure 7
<p>Effects of (<b>a</b>) pH, (<b>b</b>) initial TC concentration on removal efficiency, and (<b>c</b>) catalyst dosage.</p>
Full article ">Figure 8
<p>Free radical trapping with different free radical scavenger.</p>
Full article ">Figure 9
<p>Photocatalytic degradation mechanism of tetracycline by CNF/BiOBr.</p>
Full article ">Figure 10
<p>Possible degradation pathway of the CNF/BiOBr composite.</p>
Full article ">Figure 11
<p>Schematic diagram of the preparation route of the BiOBr/CNF.</p>
Full article ">
12 pages, 1572 KiB  
Communication
The Expression of Genes Related to Reverse Cholesterol Transport and Leptin Receptor Pathways in Peripheral Blood Mononuclear Cells Are Decreased in Morbid Obesity and Related to Liver Function
by Carlos Jiménez-Cortegana, Soledad López-Enríquez, Gonzalo Alba, Consuelo Santa-María, Gracia M. Martín-Núñez, Francisco J. Moreno-Ruiz, Sergio Valdés, Sara García-Serrano, Cristina Rodríguez-Díaz, Ailec Ho-Plágaro, María I. Fontalba-Romero, Eduardo García-Fuentes, Lourdes Garrido-Sánchez and Víctor Sánchez-Margalet
Int. J. Mol. Sci. 2024, 25(14), 7549; https://doi.org/10.3390/ijms25147549 - 9 Jul 2024
Viewed by 1279
Abstract
Obesity is frequently accompanied by non-alcoholic fatty liver disease (NAFLD). These two diseases are associated with altered lipid metabolism, in which reverse cholesterol transport (LXRα/ABCA1/ABCG1) and leptin response (leptin receptor (Ob-Rb)/Sam68) are involved. The two pathways were evaluated in peripheral blood mononuclear cells [...] Read more.
Obesity is frequently accompanied by non-alcoholic fatty liver disease (NAFLD). These two diseases are associated with altered lipid metabolism, in which reverse cholesterol transport (LXRα/ABCA1/ABCG1) and leptin response (leptin receptor (Ob-Rb)/Sam68) are involved. The two pathways were evaluated in peripheral blood mononuclear cells (PBMCs) from 86 patients with morbid obesity (MO) before and six months after Roux-en-Y gastric bypass (RYGB) and 38 non-obese subjects. In the LXRα pathway, LXRα, ABCA1, and ABCG1 mRNA expressions were decreased in MO compared to non-obese subjects (p < 0.001, respectively). Ob-Rb was decreased (p < 0.001), whereas Sam68 was increased (p < 0.001) in MO. RYGB did not change mRNA gene expressions. In the MO group, the LXRα pathway (LXRα/ABCA1/ABCG1) negatively correlated with obesity-related variables (weight, body mass index, and hip), inflammation (C-reactive protein), and liver function (alanine-aminotransferase, alkaline phosphatase, and fatty liver index), and positively with serum albumin. In the Ob-R pathway, Ob-Rb and Sam68 negatively correlated with alanine-aminotransferase and positively with albumin. The alteration of LXRα and Ob-R pathways may play an important role in NAFLD development in MO. It is possible that MO patients may require more than 6 months following RYBGB to normalize gene expression related to reverse cholesterol transport or leptin responsiveness. Full article
Show Figures

Figure 1

Figure 1
<p>LXRα, ABCA1, and ABCG1 mRNA expression in PBMCs from non-obese subjects (control) and patients with morbid obesity pre- and post-surgery. Data are represented as mean ± standard error of the mean. * <span class="html-italic">p</span> &lt; 0.001: significant differences with regard to control subjects.</p>
Full article ">Figure 2
<p>Ob-Rb and SAM68 mRNA expression in PBMCs from non-obese subjects (control) and patients with morbid obesity pre- and post-surgery. Data are represented as mean ± standard error of the mean. * <span class="html-italic">p</span> &lt; 0.001: significant differences with regard to control subjects.</p>
Full article ">Figure 3
<p>Heat maps represent those significant correlations between Ob-Rb, SAM68, LXRα, ABCA1, and ABCG1 mRNA expression in PBMCs from patients with morbid obesity pre- and post-surgery with anthropometric and biochemical variables. The Spearman correlation is displayed on a color scale from blue (positive correlation) to red (negative correlation). * Significant correlations (<span class="html-italic">p</span> &lt; 0.05). BMI: body mass index; CRP: C-reactive protein; HDL: high-density lipoprotein; HOMA-IR: homeostasis model assessment of insulin resistance; AST: aspartate aminotransferase; ALT: alanine aminotransferase; ALP: alkaline phosphatase; FLI: fatty liver index; NAFLD FS: non-alcoholic fatty liver disease fibrosis score.</p>
Full article ">
16 pages, 4110 KiB  
Article
Nanosheet BiOBr Modified Rock Wool Composites for High Efficient Oil/Water Separation and Simultaneous Dye Degradation by Activating Peroxymonosulfate
by Li Lin, Si Xiao, Chuxuan Wang, Manhong Huang, Ling Xu and Yi Huang
Molecules 2024, 29(13), 3185; https://doi.org/10.3390/molecules29133185 - 4 Jul 2024
Viewed by 1091
Abstract
The development of superlyophobic materials in liquid systems, enabling synchronous oil/water separation and dye removal from water, is highly desirable. In this study, we employed a novel superwetting array-like BiOBr nanosheets anchored on waste rock wool (RW) fibers through a simple neutralization alcoholysis [...] Read more.
The development of superlyophobic materials in liquid systems, enabling synchronous oil/water separation and dye removal from water, is highly desirable. In this study, we employed a novel superwetting array-like BiOBr nanosheets anchored on waste rock wool (RW) fibers through a simple neutralization alcoholysis method. The resulting BiOBr/RW fibers exhibited superoleophilic and superhydrophilic properties in air but demonstrated underwater superoleophobic and underoil superhydrophobic characteristics. Utilizing its dual superlyophobicity, the fiber layer demonstrated high separation efficiencies and flux velocity for oil/water mixtures by prewetting under a gravity-driven mechanism. Additionally, the novel BiOBr/RW fibers also exhibited excellent dual superlyophobicity and effective separation for immiscible oil/oil systems. Furthermore, the BiOBr/RW fibers could serve as a filter to continuously separate oil/water mixtures with high flux velocity and removal rates (>93.9%) for water-soluble dye rhodamine B (RhB) simultaneously by directly activating peroxymonosulfate (PMS) in cyclic experiments. More importantly, the mechanism of simultaneous oil/water separation and RhB degradation was proposed based on the reactive oxygen species (ROS) quenching experiments and electron paramagnetic resonance (EPR) analysis. Considering the simple modified process and the waste RW as raw material, this work may open up innovative, economical, and environmentally friendly avenues for the effective treatment of wastewater contaminated with oil and water-soluble pollutants. Full article
Show Figures

Figure 1

Figure 1
<p>Digital photos and SEM images ((<b>a</b>) pristine RW; (<b>b</b>) BiOBr/RW), EDS spectra of BiOBr/RW (<b>c</b>), and EDS mapping images of BiOBr/RW (<b>d</b>–<b>i</b>).</p>
Full article ">Figure 2
<p>XRD patterns of the samples (<b>a</b>), HRSEM images of BiOBr/RW (<b>b</b>,<b>c</b>), and HRTEM images of BiOBr/RW (<b>d</b>–<b>f</b>).</p>
Full article ">Figure 3
<p>Photos of different liquid wetting behavior and corresponding contact angle (insets) under various environments on BiOBr/RW fibers (<b>a</b>–<b>d</b>) and pristine RW (<b>e</b>–<b>h</b>), respectively.</p>
Full article ">Figure 4
<p>Photos of penetration process on BiOBr/RW surface in the air ((<b>a</b>) water droplet; (<b>b</b>): oil droplet).</p>
Full article ">Figure 5
<p>CA<sub>o/w</sub> and CA<sub>o/w</sub> of various oils detected on BiOBr/RW.</p>
Full article ">Figure 6
<p>Dynamic interaction of water droplet underoil (ligarine) and oil (1,2-dichloroethane) droplet underwater on the surface of BiOBr/RW (<b>a</b>,<b>b</b>) and pristine RW (<b>c</b>,<b>d</b>) immersed in oil or water.</p>
Full article ">Figure 7
<p>A dynamic liquid jet process on BiOBr/RW. Oil (Sudan red I dyed 1,2-dichloroethane) jet underwater (<b>a</b>,<b>b</b>) and water jet underoil (1,2-dichloroethane) (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 8
<p>Images of oil/water separation by BiOBr/RW ((<b>a</b>) prewetted by light oil; (<b>b</b>) prewetted by water; (<b>c</b>) prewetted by heavy oil; (<b>d</b>) prewetted by water); light oil and heavy oil were dyed with Sudan I.</p>
Full article ">Figure 9
<p>The separation efficiency and flux for different oil/water mixtures.</p>
Full article ">Figure 10
<p>Separation efficiency and flux of ligarine/water mixtures under different environments (<b>a</b>), the separation efficiency and contact angles detected by various immiscible liquids mixtures (Liquid 1 is diesel, n-heptane or ligarine; Liquid 2 is EG) (<b>b</b>).</p>
Full article ">Figure 11
<p>Degradation of RhB in direct PMS activation system by BiOBr/RW (<b>a</b>), and the change of UV-vis spectra with time for RhB degradation in BiOBr/RW/PMS system (<b>b</b>).</p>
Full article ">Figure 12
<p>The synchronous oil/water separation and PMS activation toward RhB removal ((<b>a</b>) BiOBr/RW; (<b>b</b>) pristine RW; (<b>c</b>) cycle results).</p>
Full article ">Figure 13
<p>Quenching experiments results in the BiOBr/RW catalyst/PMS direct activation system (<b>a</b>) and ESR spectra in different systems: (<b>b</b>) <sup>1</sup>O<sub>2</sub>, (<b>c</b>) ·SO<sub>4</sub><sup>−</sup> and ·OH, and (<b>d</b>)·O<sub>2</sub><sup>−</sup>.</p>
Full article ">Figure 14
<p>Mechanism of synchronous on-demand oil/water separation and PMS activation for pollutants removal by BiOBr/RW.</p>
Full article ">
15 pages, 5960 KiB  
Article
WO3/BiOBr S-Scheme Heterojunction Photocatalyst for Enhanced Photocatalytic CO2 Reduction
by Chen Li, Xingyu Lu, Liuyun Chen, Xinling Xie, Zuzeng Qin, Hongbing Ji and Tongming Su
Materials 2024, 17(13), 3199; https://doi.org/10.3390/ma17133199 - 30 Jun 2024
Cited by 1 | Viewed by 1029
Abstract
The photocatalytic CO2 reduction strategy driven by visible light is a practical way to solve the energy crisis. However, limited by the fast recombination of photogenerated electrons and holes in photocatalysts, photocatalytic efficiency is still low. Herein, a WO3/BiOBr S-scheme [...] Read more.
The photocatalytic CO2 reduction strategy driven by visible light is a practical way to solve the energy crisis. However, limited by the fast recombination of photogenerated electrons and holes in photocatalysts, photocatalytic efficiency is still low. Herein, a WO3/BiOBr S-scheme heterojunction was formed by combining WO3 with BiOBr, which facilitated the transfer and separation of photoinduced electrons and holes and enhanced the photocatalytic CO2 reaction. The optimized WO3/BiOBr heterostructures exhibited best activity for photocatalytic CO2 reduction without any sacrificial reagents, and the CO yield reached 17.14 μmol g−1 after reaction for 4 h, which was 1.56 times greater than that of BiOBr. The photocatalytic stability of WO3/BiOBr was also improved. Full article
(This article belongs to the Special Issue Advanced Materials for Solar Energy Utilization)
Show Figures

Figure 1

Figure 1
<p>XRD patterns (<b>a</b>), FT–IR spectra (<b>b</b>), N<sub>2</sub> adsorption–desorption isotherms (<b>c</b>), and pore size distributions (<b>d</b>) of BiOBr, WO<sub>3</sub>, and xWO<sub>3</sub>/BiOBr.</p>
Full article ">Figure 2
<p>SEM images of BiOBr (<b>a</b>), WO<sub>3</sub> (<b>b</b>), and 5WO<sub>3</sub>/BiOBr (<b>c</b>). TEM and HRTEM images (<b>d</b>–<b>f</b>), HAADF image (<b>g</b>), and corresponding EDX mapping profiles of Bi (<b>h</b>), Br (<b>i</b>), O (<b>j</b>), and W (<b>k</b>) of 5WO<sub>3</sub>/BiOBr.</p>
Full article ">Figure 3
<p>XPS spectra of Bi 4f (<b>a</b>), Br 3d (<b>b</b>), O 1 s (<b>c</b>), and W 4f (<b>d</b>) in BiOBr, WO<sub>3</sub>, and 5WO<sub>3</sub>/BiOBr.</p>
Full article ">Figure 4
<p>UV–Vis DRS of BiOBr, WO<sub>3</sub>, and xWO<sub>3</sub>/BiOBr (<b>a</b>), band gap of BiOBr and WO<sub>3</sub> (<b>b</b>), and UPS spectra (<b>c</b>) and band structure (<b>d</b>) of WO<sub>3</sub> and BiOBr.</p>
Full article ">Figure 5
<p>TRPL spectra of BiOBr and 5WO<sub>3</sub>/BiOBr (<b>a</b>), transient photocurrent density (<b>b</b>), and EIS Nyqui st plots (<b>c</b>) of BiOBr, WO<sub>3</sub>, and xWO<sub>3</sub>/BiOBr.</p>
Full article ">Figure 6
<p>Time course of photocatalytic CO<sub>2</sub> reduction over BiOBr and xWO<sub>3</sub>/BiOBr (<b>a</b>,<b>b</b>), photocatalytic CO<sub>2</sub> reduction over 5WO<sub>3</sub>/BiOBr under different conditions (<b>c</b>), and cycle test of BiOBr and 5WO<sub>3</sub>/BiOBr photocatalytic CO<sub>2</sub> reduction to CO (<b>d</b>).</p>
Full article ">Figure 7
<p>In situ DRIFTS spectra of CO<sub>2</sub> and H<sub>2</sub>O adsorption on BiOBr (<b>a</b>) and 5WO<sub>3</sub>/BiOBr (<b>b</b>) in the dark. In situ DRIFTS spectra of CO<sub>2</sub> and H<sub>2</sub>O adsorption on BiOBr (<b>c</b>) and 5WO<sub>3</sub>/BiOBr (<b>d</b>) under light irradiation.</p>
Full article ">Figure 8
<p>Band energy positions of WO<sub>3</sub> and BiOBr before (<b>a</b>) and after (<b>b</b>) contact, S-scheme charge transfer mechanism in WO<sub>3</sub>/BiOBr composites under light irradiation (<b>c</b>).</p>
Full article ">
14 pages, 6091 KiB  
Communication
A Tight-Connection g-C3N4/BiOBr (001) S-Scheme Heterojunction Photocatalyst for Boosting Photocatalytic Degradation of Organic Pollutants
by Xinyi Zhang, Weixia Li, Liangqing Hu, Mingming Gao and Jing Feng
Nanomaterials 2024, 14(13), 1071; https://doi.org/10.3390/nano14131071 - 22 Jun 2024
Cited by 1 | Viewed by 1190
Abstract
The efficient separation of photogenerated charge carriers and strong oxidizing properties can improve photocatalytic performance. Here, we combine the construction of a tightly connected S-scheme heterojunction with the exposure of an active crystal plane to prepare g-C3N4/BiOBr for [...] Read more.
The efficient separation of photogenerated charge carriers and strong oxidizing properties can improve photocatalytic performance. Here, we combine the construction of a tightly connected S-scheme heterojunction with the exposure of an active crystal plane to prepare g-C3N4/BiOBr for the degradation of high-concentration organic pollutants. This strategy effectively improves the separation efficiency of photogenerated carriers and the number of active sites. Notably, the synthesized g-C3N4/BiOBr displays excellent photocatalytic degradation activity towards various organic pollutants, including methylene blue (MB, 90.8%), congo red (CR, 99.2%), and tetracycline (TC, 89%). Furthermore, the photocatalytic degradation performance of g-C3N4/BiOBr for MB maintains 80% efficiency under natural water quality (tap water, lake water, river water), and a wide pH range (pH = 4–10). Its excellent photocatalytic activity is attributed to the tight connection between g-C3N4 and BiOBr in the S-scheme heterojunction interface, as well as the exposure of highly active (001) crystal planes. These improve the efficiency of the separation of photogenerated carriers, and maintain their strong oxidation capability. This work presents a simple approach to improving the separation of electrons and holes by tightly combining two components within a heterojunction. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns; (<b>b</b>) FT-IR spectra; and (<b>c</b>–<b>f</b>) SEM images of BiOBr, <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr.</p>
Full article ">Figure 2
<p>TEM images of (<b>a</b>) BiOBr, (<b>b</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, and (<b>c</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr; (<b>d</b>–<b>f</b>) HRTEM images of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr; (<b>g</b>–<b>l</b>) TEM mapping: Bi, Br, O, C, and N.</p>
Full article ">Figure 3
<p>(<b>a</b>) XPS spectra for the survey, (<b>b</b>) Bi 4f, (<b>c</b>) Br 3d, (<b>d</b>) O 1s, (<b>e</b>) C 1s, and (<b>f</b>) N 1s XPS spectra of BiOBr, <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV-visible diffuse-reflectance spectra; (<b>b</b>) the Tauc plot of BiOBr and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>; (<b>c</b>,<b>d</b>) the valance band XPS spectra of BiOBr and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>; and (<b>e</b>) the band structure of BiOBr and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Photocatalytic degradation efficiency of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, BiOBr, physical mixed samples, and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr; (<b>b</b>) first-order kinetic constants of MB degradation; (<b>c</b>) photocatalytic degradation of CR, TC, and MB; (<b>d</b>) initial pH (4–10); (<b>e</b>) photocatalytic degradation of MB in natural water quality; (<b>f</b>) the cyclic catalytic degradation of MB in the presence of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr.</p>
Full article ">Figure 6
<p>(<b>a</b>) PL spectra; (<b>b</b>) EIS spectra; (<b>c</b>) transient photocurrent responses; (<b>d</b>) quenching experiments.</p>
Full article ">Figure 7
<p>Photocatalytic mechanism of MB by the <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr.</p>
Full article ">
Back to TopTop