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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,787)

Search Parameters:
Keywords = plasmonics

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
8 pages, 3953 KiB  
Article
Oblique Deposited Ultra-Thin Silver Films on Polymer Gratings for Sensitive SERS Performance
by Yi-Jun Jen and Meng-Jie Lin
Nanomaterials 2024, 14(23), 1871; https://doi.org/10.3390/nano14231871 - 22 Nov 2024
Viewed by 218
Abstract
A small amount of silver was obliquely deposited onto a polymer subwavelength grating to form a metasurface that comprised silver split-tubes. An ultra-thin silver film with a monitor-controlled thickness of 20 nm at the corner of each ridge of the grating provided the [...] Read more.
A small amount of silver was obliquely deposited onto a polymer subwavelength grating to form a metasurface that comprised silver split-tubes. An ultra-thin silver film with a monitor-controlled thickness of 20 nm at the corner of each ridge of the grating provided the most sensitive surface-enhanced Raman scattering (SERS) measurements. An excitation laser beam that was incident from the substrate provided similar or better SERS enhancement than did the general configuration with the laser beam incident directly on the surface of the nanostructure. Near-field simulations were conducted to model the localized electric field enhancement and to quantify the SERS performance, demonstrating the effectiveness of this novel deposition method. Full article
Show Figures

Figure 1

Figure 1
<p>Top-view and cross-sectional SEM images of coated gratings: (<b>a</b>,<b>d</b>) GR<sub>20nm</sub><span class="html-italic">,</span> (<b>b</b>,<b>e</b>) GR<sub>80nm</sub><span class="html-italic">,</span> and (<b>c</b>,<b>f</b>) GR<sub>150nm</sub>; (<b>g</b>) schematic diagram of morphology of silver film.</p>
Full article ">Figure 2
<p>Measurement under forward and backward illumination in the transverse magnetic (TM) and transverse electric (TE) polarized state.</p>
Full article ">Figure 3
<p>TE and TM polarization spectra of reflectance (R), transmittance (T), and extinctance (E) for each sample under forward illumination.</p>
Full article ">Figure 4
<p>TE and TM polarization spectra of reflectance (R), transmittance (T), and extinctance (E) for each sample under backward illumination.</p>
Full article ">Figure 5
<p>Raman spectra of bare grating for forward and backward illumination.</p>
Full article ">Figure 6
<p>Raman spectra of GR<sub>20nm</sub>, GR<sub>80nm</sub>, and GR<sub>150nm</sub> for forward and backward illumination.</p>
Full article ">Figure 7
<p>AEF of bare grating, GR<sub>20nm</sub>, GR<sub>80nm</sub>, and GR<sub>150nm</sub> under forward illumination and backward illumination.</p>
Full article ">Figure 8
<p>Maximum steady-state amplitude of the electric field in grating substrate under forward and backward illumination.</p>
Full article ">
37 pages, 6074 KiB  
Review
Advances in Surface-Enhanced Raman Spectroscopy for Urinary Metabolite Analysis: Exploiting Noble Metal Nanohybrids
by Ningbin Zhao, Peizheng Shi, Zengxian Wang, Zhuang Sun, Kaiqiang Sun, Chen Ye, Li Fu and Cheng-Te Lin
Biosensors 2024, 14(12), 564; https://doi.org/10.3390/bios14120564 - 21 Nov 2024
Viewed by 177
Abstract
This review examines recent advances in surface-enhanced Raman spectroscopy (SERS) for urinary metabolite analysis, focusing on the development and application of noble metal nanohybrids. We explore the diverse range of hybrid materials, including carbon-based, metal–organic-framework (MOF), silicon-based, semiconductor, and polymer-based systems, which have [...] Read more.
This review examines recent advances in surface-enhanced Raman spectroscopy (SERS) for urinary metabolite analysis, focusing on the development and application of noble metal nanohybrids. We explore the diverse range of hybrid materials, including carbon-based, metal–organic-framework (MOF), silicon-based, semiconductor, and polymer-based systems, which have significantly improved SERS performance for detecting key urinary biomarkers. The principles underlying SERS enhancement in these nanohybrids are discussed, elucidating both electromagnetic and chemical enhancement mechanisms. We analyze various fabrication methods that enable precise control over nanostructure morphology, composition, and surface chemistry. The review critically evaluates the analytical performance of different hybrid systems for detecting specific urinary metabolites, considering factors such as sensitivity, selectivity, and stability. We address the analytical challenges associated with SERS-based urinary metabolite analysis, including sample preparation, matrix effects, and data interpretation. Innovative solutions, such as the integration of SERS with microfluidic devices and the application of machine learning algorithms for spectral analysis, are highlighted. The potential of these advanced SERS platforms for point-of-care diagnostics and personalized medicine is discussed, along with future perspectives on wearable SERS sensors and multi-modal analysis techniques. This comprehensive overview provides insights into the current state and future directions of SERS technology for urinary metabolite detection, emphasizing its potential to revolutionize non-invasive health monitoring and disease diagnosis. Full article
(This article belongs to the Special Issue Feature Paper in Biosensor and Bioelectronic Devices 2024)
22 pages, 5794 KiB  
Article
Hydrothermal and Co-Precipitation Combined with Photo-Reduced Preparation of Ag/AgBr/MgBi2O6 Composites for Visible Light Degradation Toward Organics
by Hsin-Yi Huang, Mudakazhi Kanakkithodi Arun, Sabu Thomas, Mei-Yao Wu, Tsunghsueh Wu and Yang-Wei Lin
Nanomaterials 2024, 14(23), 1865; https://doi.org/10.3390/nano14231865 - 21 Nov 2024
Viewed by 262
Abstract
This study developed a MgBi2O6-based photocatalyst via low-temperature hydrothermal synthesis. AgBr was co-precipitated onto MgBi2O6, and silver nanoparticles (AgNPs) were photo-reduced onto the surface. The photocatalytic performance, assessed by methylene blue (MB) degradation under white-light [...] Read more.
This study developed a MgBi2O6-based photocatalyst via low-temperature hydrothermal synthesis. AgBr was co-precipitated onto MgBi2O6, and silver nanoparticles (AgNPs) were photo-reduced onto the surface. The photocatalytic performance, assessed by methylene blue (MB) degradation under white-light LED irradiation (2.5 W, power density = 0.38 W/cm2), showed that Ag/AgBr/MgBi2O6 achieved 98.6% degradation in 40 min, outperforming MgBi2O6 (37.5%) and AgBr/MgBi2O6 (85.5%). AgNPs boosted electron-hole separation via surface plasmon resonance, reducing recombination. A Z-scheme photocatalytic mechanism was suggested, where photogenerated carriers transferred across the p–n heterojunction between AgBr and MgBi2O6, producing reactive oxygen species like superoxide and hydroxyl radicals critical for dye degradation. Thus, the Ag/AgBr/MgBi2O6 composites possessed excellent photocatalytic performance regarding dyestuff degradation (85.8–99.9% degradation within 40 min) under white-light LED irradiation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD spectra of MgBi<sub>2</sub>O<sub>6</sub>, AgBr/MgBi<sub>2</sub>O<sub>6</sub>, Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub>, and AgBr. A—AgBr, m—MgBi<sub>2</sub>O<sub>6</sub>.</p>
Full article ">Figure 2
<p>TEM images of (<b>A</b>) MgBi<sub>2</sub>O<sub>6</sub>, (<b>B</b>) AgBr/MgBi<sub>2</sub>O<sub>6</sub>, and (<b>C</b>) Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Figure 3
<p>(<b>A</b>) UV–Vis DRS spectra and (<b>B</b>) the corresponding (Ahν)<sup>2</sup> versus hν curves of MgBi<sub>2</sub>O<sub>6</sub>, AgBr/MgBi<sub>2</sub>O<sub>6</sub>, Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub>, and AgBr.</p>
Full article ">Figure 4
<p>(<b>A</b>) Photograph images and (<b>B</b>) UV–Vis spectra of MB solutions under visible light irradiation using Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites at different times.</p>
Full article ">Figure 5
<p>(<b>A</b>) Degradation performance of MB and (<b>B</b>) pseudo-first-order reaction for MB degradation by using MgBi<sub>2</sub>O<sub>6</sub>, AgBr/MgBi<sub>2</sub>O<sub>6</sub>, and Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Figure 6
<p>(<b>A</b>) PL, (<b>B</b>) Current density, and (<b>C</b>) EIS analysis of MgBi<sub>2</sub>O<sub>6</sub>, AgBr/MgBi<sub>2</sub>O<sub>6</sub>, and Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Figure 7
<p>Scavenger tests of Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Figure 8
<p>(<b>A</b>) Degradation performances of MB in the absence of catalyst and presence of P25 and Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites, (<b>B</b>) recycling used abilities, and (<b>C</b>) TOC experiments of Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Figure 9
<p>(<b>A</b>) Degradation performances of different dyestuff and photograph images of different dyestuff solutions, (<b>B</b>) degradation MB in the different environmental water samples using Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">Scheme 1
<p>Possible mechanism of photodegradation using Ag/AgBr/MgBi<sub>2</sub>O<sub>6</sub> composites.</p>
Full article ">
11 pages, 7949 KiB  
Article
Dynamic Excitation of Surface Plasmon Polaritons with Vector Laguerre–Gaussian Beams
by Aldo Peña-Ramírez, Tingting Zhai, Rafael Salas-Montiel and Víctor Ruiz-Cortés
Optics 2024, 5(4), 523-533; https://doi.org/10.3390/opt5040039 - 21 Nov 2024
Viewed by 237
Abstract
We investigate the dynamic excitation of surface plasmon polaritons (SPPs) using vector Laguerre–Gauss (LG) beams, which offer unique properties for manipulating the polarization and spatial distribution of light. Our study demonstrates the efficient coupling of SPPs with LG beams, characterized by their azimuthal [...] Read more.
We investigate the dynamic excitation of surface plasmon polaritons (SPPs) using vector Laguerre–Gauss (LG) beams, which offer unique properties for manipulating the polarization and spatial distribution of light. Our study demonstrates the efficient coupling of SPPs with LG beams, characterized by their azimuthal and radial indices (m,p), as well as polarization distribution type. Numerical simulations reveal that the vector nature of LG beams enables selective excitation of SPPs, depending on the polarization type of the beam. Experimental verification of our simulations is achieved using a gold circular Bragg grating and a spatial light modulator that generates vector LG beams. Leakage radiation imaging demonstrates the potential of vector LG beams for dynamic SPP excitation and manipulation. This study opens novel ways for the control of SPPs in plasmonic devices, such as modulators, and nanophotonic circuits. Full article
Show Figures

Figure 1

Figure 1
<p>Dynamic excitation of surface plasmon polaritons with vector Laguerre–Gaussian beams. (<b>a</b>) Schematic of the circular plasmonic Bragg grating on gold thin film. The thickness of the grating and thin film are 20 nm and 50 nm, respectively. The grating consists of 14 concentric rings with a period of a = 0.764 μm and 0.5 duty cycle. The structure is dynamically excited with vector LG beams at normal incidence. (<b>b</b>) Optical and (<b>c</b>) scanning electron microscope images of the grating. Zoom on the grating area. (<b>d</b>) Scheme of the setup for the dynamic excitation of SPPs with vector LG beams. A spatial light modulator (SLM) is used to form the vector LG beams. Red optical path shows the path for the Fourier plane.</p>
Full article ">Figure 2
<p>Calculated intensity and polarization distributions of vector LG beams <span class="html-italic">p</span> = 0 and (<b>a</b>) <span class="html-italic">m</span> = 1, types I–IV, (<b>b</b>) <span class="html-italic">m</span> = 2, types I–IV, (<b>c</b>) <span class="html-italic">m</span> = 3, types I–IV, and (<b>d</b>) <span class="html-italic">m</span> = 4, types I–IV. Scale bar is 25 μm.</p>
Full article ">Figure 3
<p>Excitation of surface plasmon polaritons with type I–IV vector <math display="inline"><semantics> <mrow> <mi>L</mi> <msub> <mi>G</mi> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mo>[</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> <mo>]</mo> <mo>)</mo> </mrow> </msub> </mrow> </semantics></math> beams. Distribution of the electric field and magnetic field lines for vector <math display="inline"><semantics> <mrow> <mi>L</mi> <msub> <mi>G</mi> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mi>m</mi> <mo>)</mo> </mrow> </msub> </mrow> </semantics></math> beams (<b>a</b>) <span class="html-italic">m</span> = 1, type I, (<b>b</b>) <span class="html-italic">m</span> = 1, type II, (<b>c</b>) <span class="html-italic">m</span> = 1, type III, and (<b>d</b>) <span class="html-italic">m</span> = 1, type IV, (<b>e</b>) <span class="html-italic">m</span> = 2, type I, (<b>f</b>) <span class="html-italic">m</span> = 2, type II, (<b>g</b>) <span class="html-italic">m</span> = 2, type III, and (<b>h</b>) <span class="html-italic">m</span> = 2, type IV, (<b>i</b>) <span class="html-italic">m</span> = 3, type I, (<b>j</b>) <span class="html-italic">m</span> = 3, type II, (<b>k</b>) <span class="html-italic">m</span> = 3, type III, and (<b>l</b>) <span class="html-italic">m</span> = 3, type IV, (<b>m</b>) <span class="html-italic">m</span> = 4, type I, (<b>n</b>) <span class="html-italic">m</span> = 4, type II, (<b>o</b>) <span class="html-italic">m</span> = 4, type III, and (<b>p</b>) <span class="html-italic">m</span> = 4, type IV. Scale bar is 500 nm.</p>
Full article ">Figure 4
<p>Experimental excitation of surface plasmon polaritons. (<b>a</b>) Optical image of the gold Bragg grating with a period of 764 nm. The two circumferences represent the inner and outer ring of the grating. Leakage radiation imaging of the excitation of SPPs with vector <math display="inline"><semantics> <msub> <mrow> <mi>LG</mi> </mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </msub> </semantics></math> (<b>b</b>) type III and (<b>c</b>) type I beams. Only type III beams excite SPPs.</p>
Full article ">Figure 5
<p>Experimental excitation of surface plasmon polaritons. Type I and III vector <math display="inline"><semantics> <msub> <mrow> <mi>LG</mi> </mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </msub> </semantics></math> beams. Experimental distribution of polarization of the incident field types (<b>a</b>) I and (<b>d</b>) III. Leakage radiation microscopy images of the excitation of SPPs in (<b>b</b>) the direct and (<b>c</b>) Fourier spaces for the type I and (<b>e</b>,<b>f</b>) for type III beams. Scheme of the incident fields for simulations of type (<b>g</b>) I and (<b>j</b>) III beams. Corresponding results of the excitation of SPPs in the (<b>h</b>,<b>k</b>) direct and (<b>i</b>,<b>l</b>) Fourier spaces.</p>
Full article ">Figure 6
<p>Excitation of SPPs with combinations of vector <math display="inline"><semantics> <msub> <mrow> <mi>LG</mi> </mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </msub> </semantics></math> type I and III beams. Scheme of distributions (<b>a</b>) D1, (<b>b</b>) D2, (<b>c</b>) D3, and (<b>d</b>) D4. The white lines represent the inner and outer grating rings. Leakage radiation images of SPPs in (<b>e</b>–<b>h</b>) the direct and (<b>i</b>–<b>l</b>) Fourier spaces for distributions D1, D2, D3, and D4, respectively. Arrows indicate the sections where leakage radiation was detected. Insets in (<b>e</b>–<b>h</b>) are magnifications in the center of the grating. Scale bar is 10 μm in (<b>a</b>–<b>h</b>).</p>
Full article ">
36 pages, 9567 KiB  
Review
Two-Dimensional MoS2-Based Photodetectors
by Leilei Ye, Xiaorong Gan and Romana Schirhagl
Sustainability 2024, 16(22), 10137; https://doi.org/10.3390/su162210137 - 20 Nov 2024
Viewed by 203
Abstract
Nanomaterials can significantly improve the analytical performance of optical sensors for environmental pollutants. Two-dimensional (2D) molybdenum sulfide (MoS2) exhibits some unique physicochemical properties, such as strong light–matter interactions, bandgap tunability, and high carrier mobility, which are beneficial for constructing flexible optoelectronic [...] Read more.
Nanomaterials can significantly improve the analytical performance of optical sensors for environmental pollutants. Two-dimensional (2D) molybdenum sulfide (MoS2) exhibits some unique physicochemical properties, such as strong light–matter interactions, bandgap tunability, and high carrier mobility, which are beneficial for constructing flexible optoelectronic devices. In this review, the principle and classification of 2D MoS2-based photodetectors (PDs) are introduced, followed by a discussion about the physicochemical properties of 2D MoS2, as well as the structure–property relationships of 2D MoS2-based photoactive materials for PDs to understand the modulation strategies for enhancing the photodetection performance. Furthermore, we discuss significant advances in the surface modification and functionalization of 2D MoS2 for developing high-performance PDs, particularly focusing on synthesis pathways, modification strategies, and underlying physiochemical mechanisms for enhanced photodetection capabilities. Finally, conclusions and research perspectives on resolving significant bottlenecks or remaining challenges are offered based on recent developments in 2D MoS2-based PDs. Full article
Show Figures

Figure 1

Figure 1
<p>Challenges of using pristine 2D MoS<sub>2</sub> in PDs and typical strategies for improving photodetection performances.</p>
Full article ">Figure 2
<p>Metal plasmonic structure dynamics. (<b>a</b>) Electron behavior in a plasmonic NP under solar illumination. (<b>b</b>) A mechanical harmonic oscillator is used to explain the coherent electronic cloud on the metal surface under light forces. Reproduced with permission from ref. [<a href="#B74-sustainability-16-10137" class="html-bibr">74</a>]. Copyright 2019, American Chemical Society.</p>
Full article ">Figure 3
<p>Schematic of three typical optical resonance cavities. (<b>a</b>) Fabry–Perot-type cavity. (<b>b</b>) The 2D photonic crystal cavity. (<b>c</b>) Whispering gallery mode microcavity. Reproduced with permission from ref. [<a href="#B92-sustainability-16-10137" class="html-bibr">92</a>]. Copyright 2020 John Wiley &amp; Sons, Inc.</p>
Full article ">Figure 4
<p>(<b>a</b>) Photoresponse of defective-MoS<sub>2</sub> PDs under vacuum or when exposed to air. (<b>b</b>) Gas response of a defective MoS<sub>2</sub> PD with and without MA<sub>3</sub>Bi<sub>2</sub>Br<sub>9</sub> treatment. (<b>c</b>) The schematic illustration of defect states in the electronic band of MoS<sub>2</sub>. Reproduced with permission from ref. [<a href="#B103-sustainability-16-10137" class="html-bibr">103</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 5
<p>Band structures of bulk MoS<sub>2</sub> (<b>a</b>), quadrilayer MoS<sub>2</sub> (<b>b</b>), bilayer MoS<sub>2</sub> (<b>c</b>), and monolayer MoS<sub>2</sub> (<b>d</b>). Reproduced with permission from ref. [<a href="#B116-sustainability-16-10137" class="html-bibr">116</a>]. Copyright 2010 American Chemical Society.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic presentation of the synthesis processes of wrinkly MoS<sub>2</sub> flakes. (<b>b</b>) Optical microscopy image and (<b>c</b>) atomic force microscopy image of a wrinkled MoS<sub>2</sub> flake (<b>d</b>) Photoluminescence spectra of wrinkly MoS<sub>2</sub> flakes at different locations (flat and wrinkly parts). (<b>e</b>) Relationship between the direct bandgap transition and the localized tensile strain. Reproduced with permission from ref. [<a href="#B114-sustainability-16-10137" class="html-bibr">114</a>]. Copyright 2013 American Chemical Society.</p>
Full article ">Figure 7
<p>(<b>a</b>) The relationship between the S/Mo ratio (<span class="html-italic">x</span>) and bandgaps of 2D MoS<span class="html-italic"><sub>x</sub></span>. (<b>b</b>) Interface energy band diagram between Au electrodes and 2D MoS<span class="html-italic"><sub>x</sub></span> before (<b>b</b>) and after (<b>c</b>) applying THz radiation. Electron-transport behavior in PDs based on 2D MoS<span class="html-italic"><sub>x</sub></span> with (<b>d</b>) and without (<b>e</b>) applying THz radiation. (<b>f</b>) The relationship between the photocurrent or <b>R</b> and bias voltage. Reproduced with permission from ref. [<a href="#B141-sustainability-16-10137" class="html-bibr">141</a>]. Copyright 2020 American Chemical Society.</p>
Full article ">Figure 8
<p>(<b>a</b>) A diagram showing the setup utilized in an MoS<sub>2</sub> PD. (<b>b</b>) The relationship between strain and temperature of polycarbonate. (<b>c</b>) Differential reflectance spectra measured at different temperatures. Reproduced with permission from ref. [<a href="#B149-sustainability-16-10137" class="html-bibr">149</a>]. Copyright 2019 Elsevier.</p>
Full article ">Figure 9
<p>Graphical illustration of mixed-dimensional heterostructures with different combination formats including (<b>a</b>) 0D/2D, (<b>b</b>) 1D/2D, and (<b>c</b>) 2D/3D mix-dimensional structures.</p>
Full article ">Figure 10
<p>(<b>a</b>) Schematic presentation of LSPR in Au-NP grating. (<b>b</b>) Extinction spectra of different Au NPs and (<b>c</b>) transfer characteristics of the Au-NP/MoS<sub>2</sub> PD. Reproduced with permission from ref. [<a href="#B170-sustainability-16-10137" class="html-bibr">170</a>]. Copyright 2020 American Chemical Society. (<b>d</b>) Schematic diagrams of Au-MoS<sub>2</sub>-Au PDs. (<b>e</b>) Light intensity-dependent <b>R</b> of Au-MoS<sub>2</sub> and Au-MoS<sub>2</sub>-Au PDs. (<b>f</b>) Schematic presentation of the interface energy band of Au-MoS<sub>2</sub>-Au. (<b>g</b>) Calculated electrical field strength distributions of Au NPs on the two surfaces of MoS<sub>2</sub>. (<b>h</b>) Photocurrents of Au-MoS<sub>2</sub>-Au PDs under different light intensities. (<b>i</b>) Possible charge-transfer processes of Au-MoS<sub>2</sub>-Au PDs under visible light. Reproduced with permission from ref. [<a href="#B171-sustainability-16-10137" class="html-bibr">171</a>]. Copyright 2022 American Chemical Society.</p>
Full article ">Figure 11
<p>(<b>a</b>) Schematic representation of MoS<sub>2</sub> PD (MNPs refer to the MXene QDs with an average size of 8 nm). (<b>b</b>) I-V plots of pristine MoS<sub>2</sub> and MXene-QD/MoS<sub>2</sub> PD in the dark. Two-dimensional photoresponse maps of (<b>c</b>) MXene-QD/MoS<sub>2</sub> PDs and (<b>d</b>) pristine MoS<sub>2</sub> under 635 nm excitation wavelengths. Reproduced with permission from ref. [<a href="#B176-sustainability-16-10137" class="html-bibr">176</a>]. Copyright 2022 American Chemical Society.</p>
Full article ">Figure 12
<p>(<b>a</b>) A 3D schematic representation of SnS<sub>2</sub>-QDs/MoS<sub>2</sub> PD. (<b>b</b>) Schematic illustration of band structures of SnS<sub>2</sub> QDs and monolayer MoS<sub>2</sub>. (<b>c</b>) <b>R</b> of SnS<sub>2</sub>-QDs/MoS<sub>2</sub> PD. Reproduced with permission from ref. [<a href="#B174-sustainability-16-10137" class="html-bibr">174</a>]. Copyright 2022 American Chemical Society.</p>
Full article ">Figure 13
<p>Device structure and interfacial charge-carrier transport behavior for n-ZnO/p-MoS<sub>2</sub> PDs with a forward biasing voltage (<b>a</b>), with a forward biasing voltage and under UV light (<b>b</b>), and with a forward biasing voltage and under green light (<b>c</b>). Reproduced with permission from ref. [<a href="#B182-sustainability-16-10137" class="html-bibr">182</a>]. Copyright 2020 John Wiley &amp; Sons, Inc.</p>
Full article ">Figure 14
<p>Schematic presentation of the device structure (<b>a</b>), the plot of <b>R</b> vs. wavelength (<b>b</b>), and the plot of current vs. decay time (<b>c</b>) for the 2D-MoS<sub>2</sub>/1D-CuO heterojunction PD. Reproduced with permission from ref. [<a href="#B184-sustainability-16-10137" class="html-bibr">184</a>]. Copyright 2016 American Chemical Society. (<b>d</b>) Schematic diagram of the PD device based on vertical MoS<sub>2</sub> nanosheets/p-GaN NRs. (<b>e</b>) SEM image of vertical MoS<sub>2</sub> nanosheets. (<b>f</b>) The dark and illuminated I-V curves of the PD device. Reproduced with permission from ref. [<a href="#B187-sustainability-16-10137" class="html-bibr">187</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 15
<p>(<b>a</b>) Schematic presentation of the device structure of an MoS<sub>2</sub>/CsPbBr<sub>3</sub> PD. Charge production and transport processes at the MoS<sub>2</sub>/CsPbBr<sub>3</sub> heterojunction in the dark (<b>b</b>) and under light illumination (<b>c</b>). (<b>d</b>) UV-vis absorption spectra of the MoS<sub>2</sub>, CsPbBr<sub>3</sub>, and MoS<sub>2</sub>/CsPbBr<sub>3</sub> heterojunction. (<b>e</b>) I–V characteristics plots of the MoS<sub>2</sub>, CsPbBr<sub>3</sub>, and MoS<sub>2</sub>/CsPbBr<sub>3</sub> heterojunction with and without laser illumination. (<b>f</b>) Relationship between <b>R</b> and power intensity. Reproduced with permission from ref. [<a href="#B191-sustainability-16-10137" class="html-bibr">191</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 16
<p>(<b>a</b>) Schematic presentation of the device structure of graphene/2D MoS<sub>2</sub> PD under illumination. (<b>b</b>) A diagram of the photocurrent generation of graphene/2D MoS<sub>2</sub> PD. (<b>c</b>) Current–voltage curve of graphene/2D MoS<sub>2</sub> PD in the dark and under 100 mW/cm<sup>2</sup> illumination. (<b>d</b>) <b>R</b> of graphene/2D MoS<sub>2</sub> PD at different wavelengths. (<b>e</b>) Photoconductivity of graphene. Reproduced with permission from ref. [<a href="#B194-sustainability-16-10137" class="html-bibr">194</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 17
<p>(<b>a</b>) Schematic picture of a CZTS/MoS<sub>2</sub> PD. (<b>b</b>) Charge transfer mechanism at CZTS/MoS<sub>2</sub> heterojunctions. (<b>c</b>) Estimated <b>R</b> of the CZTS/MoS<sub>2</sub> heterojunction PD for various wavelengths at a bias of 6 V. Reproduced with permission from ref. [<a href="#B215-sustainability-16-10137" class="html-bibr">215</a>]. Copyright 2020 Elsevier.</p>
Full article ">Figure 18
<p>(<b>a</b>) Schematic presentation of 3D structures of MoS<sub>2</sub>/MoSe<sub>2</sub>/GaN heterostructures. (<b>b</b>) Schematics of the energy band diagram for the heterostructures. (<b>c</b>) Photoswitching of D<sub>1</sub>, D<sub>2</sub>, D<sub>3</sub>, and D<sub>4</sub> devices (D<sub>1</sub>, GaN/MoSe<sub>2</sub>; D<sub>2</sub>, GaN/MoS<sub>2</sub>; D<sub>3</sub>, GaN/MoS<sub>2</sub>/MoSe<sub>2</sub>; D<sub>4</sub>, GaN/MoSe<sub>2</sub>/MoS<sub>2</sub>). (<b>d</b>) UV (λ = 365 nm) intensity-dependent <b>R</b> for different device configurations. (<b>e</b>) Specific detectivity for different device configurations. Reproduced with permission from ref. [<a href="#B220-sustainability-16-10137" class="html-bibr">220</a>]. Copyright 2023 American Chemical Society.</p>
Full article ">
56 pages, 5775 KiB  
Review
Gold Nanoparticles in Nanomedicine: Unique Properties and Therapeutic Potential
by Furkan Eker, Emir Akdaşçi, Hatice Duman, Mikhael Bechelany and Sercan Karav
Nanomaterials 2024, 14(22), 1854; https://doi.org/10.3390/nano14221854 - 20 Nov 2024
Viewed by 574
Abstract
Gold nanoparticles (NPs) have demonstrated significance in several important fields, including drug delivery and anticancer research, due to their unique properties. Gold NPs possess significant optical characteristics that enhance their application in biosensor development for diagnosis, in photothermal and photodynamic therapies for anticancer [...] Read more.
Gold nanoparticles (NPs) have demonstrated significance in several important fields, including drug delivery and anticancer research, due to their unique properties. Gold NPs possess significant optical characteristics that enhance their application in biosensor development for diagnosis, in photothermal and photodynamic therapies for anticancer treatment, and in targeted drug delivery and bioimaging. The broad surface modification possibilities of gold NPs have been utilized in the delivery of various molecules, including nucleic acids, drugs, and proteins. Moreover, gold NPs possess strong localized surface plasmon resonance (LSPR) properties, facilitating their use in surface-enhanced Raman scattering for precise and efficient biomolecule detection. These optical properties are extensively utilized in anticancer research. Both photothermal and photodynamic therapies show significant results in anticancer treatments using gold NPs. Additionally, the properties of gold NPs demonstrate potential in other biological areas, particularly in antimicrobial activity. In addition to delivering antigens, peptides, and antibiotics to enhance antimicrobial activity, gold NPs can penetrate cell membranes and induce apoptosis through various intracellular mechanisms. Among other types of metal NPs, gold NPs show more tolerable toxicity capacity, supporting their application in wide-ranging areas. Gold NPs hold a special position in nanomaterial research, offering limited toxicity and unique properties. This review aims to address recently highlighted applications and the current status of gold NP research and to discuss their future in nanomedicine. Full article
Show Figures

Figure 1

Figure 1
<p>Applications of gold NPs in various fields [<a href="#B3-nanomaterials-14-01854" class="html-bibr">3</a>,<a href="#B5-nanomaterials-14-01854" class="html-bibr">5</a>].</p>
Full article ">Figure 2
<p>Graph representing the published research papers that include “gold nanoparticles” in their title for the last 5 years, with a pie chart showing the distribution of applications based on the discussed sections [<a href="#B8-nanomaterials-14-01854" class="html-bibr">8</a>].</p>
Full article ">Figure 3
<p>General properties of gold nanoparticles [<a href="#B13-nanomaterials-14-01854" class="html-bibr">13</a>].</p>
Full article ">Figure 4
<p>Drug delivery application of gold NPs [<a href="#B29-nanomaterials-14-01854" class="html-bibr">29</a>].</p>
Full article ">Figure 5
<p>Nucleic acid delivery mechanism of gold NPs. Through endocytosis, functionalized gold NPs effectively transport nucleic acids into cells, and surface alterations improve targeting. The nucleic acids are released into the cytoplasm by endosomal escape mechanisms after internalization, providing opportunities for immunotherapy and gene therapy [<a href="#B61-nanomaterials-14-01854" class="html-bibr">61</a>].</p>
Full article ">Figure 6
<p>Representation of protein delivery mechanism of gold NPs. By altering their surfaces with ligands, polymers, or linkers, gold NPs may be made to bind particular proteins. This increases their circu-lation time and stops enzymatic breakdown. Through endocytosis, gold NPs enable cellular ab-sorption and release protein cargo inside cells. Therapeutic applications benefit from surface changes that improve targeting to certain tissues or cell types [<a href="#B75-nanomaterials-14-01854" class="html-bibr">75</a>].</p>
Full article ">Figure 7
<p>Gold NP-based photothermal and photodynamic therapy in anticancer application [<a href="#B52-nanomaterials-14-01854" class="html-bibr">52</a>,<a href="#B129-nanomaterials-14-01854" class="html-bibr">129</a>].</p>
Full article ">Figure 8
<p>Antibacterial activity of gold NPs by multiple mechanisms [<a href="#B209-nanomaterials-14-01854" class="html-bibr">209</a>].</p>
Full article ">Figure 9
<p>Potential toxicity mechanisms of gold NPs [<a href="#B257-nanomaterials-14-01854" class="html-bibr">257</a>,<a href="#B258-nanomaterials-14-01854" class="html-bibr">258</a>].</p>
Full article ">Figure 10
<p>Number of registered patents containing “Gold Nanoparticle” in their title in the last five years [<a href="#B277-nanomaterials-14-01854" class="html-bibr">277</a>].</p>
Full article ">
15 pages, 21984 KiB  
Article
Green Synthesis and Characterization of Silver Nanoparticles from Minthostachys acris Schmidt Lebuhn (Muña) and Its Evaluation as a Bactericidal Agent Against Escherichia coli and Staphylococus aureus
by Fabián Ccahuana Ayma, Ana María Osorio Anaya, Gabrielle Caroline Peiter, Silvia Jaerger and Ricardo Schneider
Micro 2024, 4(4), 706-720; https://doi.org/10.3390/micro4040043 - 20 Nov 2024
Viewed by 410
Abstract
The search for new synthesis methodologies based on the principles of green chemistry has led to various studies for the production of silver nanoparticles (AgNPs) using extracts from different parts of plants. Based on this, the present study aims to carry out green [...] Read more.
The search for new synthesis methodologies based on the principles of green chemistry has led to various studies for the production of silver nanoparticles (AgNPs) using extracts from different parts of plants. Based on this, the present study aims to carry out green synthesis (biosynthesis), characterization, and antibacterial evaluation of reduced and stabilized silver nanoparticles (AgNPs) with aqueous extracts of Minthostachys acris in a simple, ecological, and environmentally safe manner. The extraction process of the organic components is performed using two methods: immersion and the agitation of the leaves of Minthostachys acris Schmidt Lebuhn (Muña) at 0.1% for different times (0.5, 1, 3, 6, and 10 min). Compounds such as hydroxycinnamic acid derivatives, quinic, caffeic, rosmarinic acids, and flavonols present in the Muña extract facilitate the formation of AgNPs; this compounds act as a coating and stabilizing agent. The bioactive components from natural resources facilitate the formation of AgNPs, partially or completely replacing the contaminating and toxic elements present in chemical reagents. The biosynthesis is carried out at room temperature for pH 7 and 8. The synthesized AgNPs are characterized by UV-visible spectroscopy to identify the surface plasmon resonance (SPR) band, which shows an absorption peak around 419 nm and 423 nm for pH 7 and p.H 8, respectively, and Fourier-transform infrared spectroscopy (FTIR) to identify the possible biomolecules responsible for bioreduction and stabilization, with a peak at 1634 cm−1. Dynamic light scattering (DLS) shows the hydrodynamic size of the colloidal nanoparticles between 11 and 200 nm, and scanning electron microscopy (SEM) reveals monodisperse AgNPs of different morphologies, mostly nanospheres, while Laser-Induced Breakdown Spectroscopy (LIBS) demonstrates the presence of Ag in the colloidal solution. The evaluation of the bactericidal activity of the AgNPs using the disk diffusion method against Escherichia coli (E. coli) and Staphylococus aureus (S.aureus) shows that the synthesized AgNPs have effective antibacterial activity against E. coli for the extracts obtained at 6 min for both the immersion and agitation methods, respectively. The significance of this work lies in the use of bioactive components from plants to obtain AgNPs in a simple, rapid, and economical way, with potential applications in biomedical fields. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Minthostachys acris</span> Schmidt Lebuhn.</p>
Full article ">Figure 2
<p>UV-vis spectra of obtained AgNPs at pH 7 (<b>a</b>) by the immersion method and (<b>b</b>) by the stirring method. The time is refereed to the extraction time.</p>
Full article ">Figure 3
<p>UV-vis spectra of AgNPs at pH 8 (<b>a</b>) by the immersion method and (<b>b</b>) by the stirring method.</p>
Full article ">Figure 4
<p>Proposed synthesis of AgNPs using <span class="html-italic">M. acris</span> extract with silver nitrate (AgNO<sub>3</sub>) solution.</p>
Full article ">Figure 5
<p>FT−IR Spectra by the immersion method of (<b>a</b>) <span class="html-italic">M. acris</span> extract, (<b>b</b>) AgNPs at pH 7, and (<b>c</b>) at pH 8.</p>
Full article ">Figure 6
<p>FT−IR Spectra by the stirring method of (<b>a</b>) <span class="html-italic">M. acris</span> extract, (<b>b</b>) AgNPs at pH 7, and (<b>c</b>) at pH 8.</p>
Full article ">Figure 7
<p>Dynamic laser scattering for AgNPs obtained at pH 7 and 10 min.</p>
Full article ">Figure 8
<p>Morphological study of silver nanoparticles by scanning electron microscopy of AgNPs synthesized at pH 7 (<b>a</b>) by the immersion method for 10 min and (<b>b</b>) by the stirring method for 10 min.</p>
Full article ">Figure 9
<p>LIBS of aqueous solution and AgNPs supported in filter paper.</p>
Full article ">Figure 10
<p>Quantitative evaluation of antibacterial activity of AgNPs against <span class="html-italic">Escherichia coli</span>.</p>
Full article ">Figure 11
<p>Quantitative evaluation of antibacterial activity of AgNPs against <span class="html-italic">Staphylococus aureus</span>.</p>
Full article ">
19 pages, 5144 KiB  
Article
An Optimized Graphene-Based Surface Plasmon Resonance Biosensor for Detecting SARS-CoV-2
by Talia Tene, Fabian Arias Arias, Karina I. Paredes-Páliz, Camilo Haro-Barroso and Cristian Vacacela Gomez
Appl. Sci. 2024, 14(22), 10724; https://doi.org/10.3390/app142210724 - 19 Nov 2024
Viewed by 326
Abstract
Graphene-enhanced surface plasmon resonance (SPR) biosensors offer promising advancements in viral detection, particularly for SARS-CoV-2. This study presents the design and optimization of a multilayer SPR biosensor incorporating silver, silicon nitride, single-layer graphene, and thiol-tethered ssDNA to achieve high sensitivity and specificity. Key [...] Read more.
Graphene-enhanced surface plasmon resonance (SPR) biosensors offer promising advancements in viral detection, particularly for SARS-CoV-2. This study presents the design and optimization of a multilayer SPR biosensor incorporating silver, silicon nitride, single-layer graphene, and thiol-tethered ssDNA to achieve high sensitivity and specificity. Key metrics, including SPR angle shift (Δθ), sensitivity (S), detection accuracy (DA), and figure of merit (FoM), were assessed across SARS-CoV-2 concentrations from 150 to 525 mM. The optimized biosensor achieved a sensitivity of 315.91°/RIU at 275 mM and a maximum Δθ of 4.2° at 400 mM, demonstrating strong responsiveness to virus binding. The sensor maintained optimal accuracy and figure of merit at lower concentrations, with a linear sensitivity response up to 400 mM, after which surface saturation limited further responsiveness. These results highlight the suitability of the optimized biosensor for real-time, point-of-care SARS-CoV-2 detection, particularly at low viral loads, supporting its potential in early diagnostics and epidemiological monitoring. Full article
(This article belongs to the Special Issue Advanced Photonic Metamaterials and Its Applications)
Show Figures

Figure 1

Figure 1
<p>Proposed SPR sensors for SARS-CoV-2 sensing.</p>
Full article ">Figure 2
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 60° to 80° for the systems Sys<sub>0</sub> to Sys<sub>9</sub>. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage of all systems under consideration.</p>
Full article ">Figure 3
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 60° to 80°, considering different silver thicknesses from 40 to 65 nm. Ag<sub>base</sub> denotes Sys<sub>8</sub> in water with an initial silver thickness of 55 nm. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage for all silver thicknesses tested.</p>
Full article ">Figure 4
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 65° to 90°, considering different silicon nitride thicknesses from 5 to 20 nm. S<sub>3</sub>N<sub>4_base</sub> denotes Sys<sub>8</sub> in water with the optimized silver thickness of 50 nm and initial silicon nitride thickness of 5 nm. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage for all silicon nitride thickness tested.</p>
Full article ">Figure 5
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 65° to 90°, increasing the number of graphene layers. L1<sub>base</sub> denotes Sys<sub>8</sub> in water with the optimized silver thickness of 50 nm, the optimized silicon nitride thickness of 15 nm, and the initial one-layer graphene system. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage for all number of graphene layers tested.</p>
Full article ">Figure 6
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 70° to 80°, increasing the thickness of ssDNA layers. ssDNA<sub>3.2_base</sub> denotes Sys<sub>8</sub> in water with the optimized silver thickness of 50 nm, the optimized silicon nitride thickness of 15 nm, the optimized single-layer graphene, and the initial ssDNA thickness of 3.2 nm. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage for all ssDNA thickness tested.</p>
Full article ">Figure 7
<p>(<b>a</b>) SPR curves as a function of the angle of incidence ranging from 65° to 90°, considering different virus concentrations from 150 to 525 mM. The black curve (n<sub>1.334</sub>) denotes the optimized Sys<sub>8</sub> in PBS solution before virus adsorption. (<b>b</b>) Attenuation percentage, (<b>c</b>) FWHM, and (<b>d</b>) sensitivity enhancement percentage for all virus concentrations tested.</p>
Full article ">Figure 8
<p>(<b>a</b>) Variation angle after the adsorption of SARS-CoV-2 at different concentrations from 0 to 525 mM. Performance metrics: (<b>b</b>) sensitivity, (<b>c</b>) detection accuracy, and (<b>d</b>) figure of merit as a function of the SARS-CoV-2 concentration.</p>
Full article ">
28 pages, 7444 KiB  
Article
Exploring the Potential of Biomimetic Peptides in Targeting Fibrillar and Filamentous Alpha-Synuclein—An In Silico and Experimental Approach to Parkinson’s Disease
by Sophia A. Frantzeskos, Mary A. Biggs and Ipsita A. Banerjee
Biomimetics 2024, 9(11), 705; https://doi.org/10.3390/biomimetics9110705 - 18 Nov 2024
Viewed by 576
Abstract
Alpha-synuclein (ASyn) is a protein that is known to play a critical role in Parkinson’s disease (PD) due to its propensity for misfolding and aggregation. Furthermore, this process leads to oxidative stress and the formation of free radicals that cause neuronal damage. In [...] Read more.
Alpha-synuclein (ASyn) is a protein that is known to play a critical role in Parkinson’s disease (PD) due to its propensity for misfolding and aggregation. Furthermore, this process leads to oxidative stress and the formation of free radicals that cause neuronal damage. In this study, we have utilized a biomimetic approach to design new peptides derived from marine natural resources. The peptides were designed using a peptide scrambling approach where antioxidant moieties were combined with fibrillary inhibition motifs in order to design peptides that would have a dual targeting effect on ASyn misfolding. Of the 20 designed peptides, 12 were selected for examining binding interactions through molecular docking and molecular dynamics approaches, which revealed that the peptides were binding to the pre-NAC and NAC (non-amyloid component) domain residues such as Tyr39, Asn65, Gly86, and Ala85, among others. Because ASyn filaments derived from Lewy body dementia (LBD) have a different secondary structure compared to pathogenic ASyn fibrils, both forms were tested computationally. Five of those peptides were utilized for laboratory validation based on those results. The binding interactions with fibrils were confirmed using surface plasmon resonance studies, where EQALMPWIWYWKDPNGS, PYYYWKDPNGS, and PYYYWKELAQM showed higher binding. Secondary structural analyses revealed their ability to induce conformational changes in ASyn fibrils. Additionally, PYYYWKDPNGS and PYYYWKELAQM also demonstrated antioxidant properties. This study provides insight into the binding interactions of varying forms of ASyn implicated in PD. The peptides may be further investigated for mitigating fibrillation at the cellular level and may have the potential to target ASyn. Full article
Show Figures

Figure 1

Figure 1
<p>SiteMap analysis showing binding pocket regions in (<b>a</b>) Lewy body dementia-derived alpha-synuclein filament and (<b>b</b>) pathogenic alpha-synuclein fibrils. The explicit regions within the binding pocket are color-coded as follows: Hydrophilic regions—green; hydrophobic—yellow; hydrogen bond donor region—blue; H-bond acceptor region—red.</p>
Full article ">Figure 2
<p>PLIP analysis for peptides with alpha-synuclein filaments derived from Lewy body dementia. (<b>a</b>) PYYYWKDPNGS; (<b>b</b>) PIWWYWKDPNGS; (<b>c</b>) PYYYWKELAQM; (<b>d</b>) PIWWYWKELAQM; (<b>e</b>) PWIWYWKDPNGS; (<b>f</b>) EQALMPWIWYWKDPNGS; (<b>g</b>) ELAQMPYYYWKDPNG; (<b>h</b>) ELAQMPIWWYWKDPNGS; (<b>i</b>) DPNGSPYYYWKELAQM; (<b>j</b>) DPNGSPIWWYWKELAQM; (<b>k</b>) ELAQMGPEGPMGLEDPNGS; (<b>l</b>) EQALMGFYGPTEDPNGS.</p>
Full article ">Figure 3
<p>PLIP analysis for peptides with normal alpha-synuclein fibrils. (<b>a</b>) PYYYWKDPNGS; (<b>b</b>) PIWWYWKDPNGS; (<b>c</b>) PYYYWKELAQM; (<b>d</b>) PIWWYWKELAQM; (<b>e</b>) PWIWYWKDPNGS; (<b>f</b>) EQALMPWIWYWKDPNGS; (<b>g</b>) ELAQMPYYYWKDPNG; (<b>h</b>) ELAQMPIWWYWKDPNGS; (<b>i</b>) DPNGSPYYYWKELAQM; (<b>j</b>) DPNGSPIWWYWKELAQM; (<b>k</b>) ELAQMGPEGPMGLEDPNGS; (<b>l</b>) EQALMGFYGPTEDPNGS.</p>
Full article ">Figure 4
<p>Trajectory images over 100 ns MD simulations with ASyn LBD-derived filament with (<b>a</b>) EQALMPWIWYWKDPNGS (green), (<b>b</b>) ELAQMGPEGPMGLEDPNGS (dark blue), (<b>c</b>) PYYYWKDPNGS (light blue), and (<b>d</b>) PIWWYWKELAQM (brown).</p>
Full article ">Figure 5
<p>Trajectory images over 100 ns MD simulations with ASyn pathogenic fibrils. (<b>a</b>) EQALMGFYGPTEDPNGS (black); (<b>b</b>) DPNGSPYYYWKELAQM (yellow); (<b>c</b>) PYYYWKELAQM (green); (<b>d</b>) PIWWYWKELAQM (purple).</p>
Full article ">Figure 6
<p>Comparison of root mean square fluctuation (RMSF) plots of the designed peptides with the (<b>a</b>) ASyn filament derived from Lewy body dementia brains and (<b>b</b>) pathogenic ASyn fibrils.</p>
Full article ">Figure 7
<p>SPR sensograms showing binding with ASyn aggregates. Top row, from left to right: PYYYWKDPNGS; PYYYWKELAQM and EQALMPWIWYWKDPNGS. Bottom row, from left to right: ELAQPYYYWKDPNGS and ELAQPEGPMGLEDPNGS.</p>
Full article ">Figure 8
<p>Comparison of fluorescence of thioflavin T over time for ASyn fibrils before (control) and after incubation with peptides.</p>
Full article ">Figure 9
<p>Comparison of the secondary structures of ASyn upon incubation with peptides.</p>
Full article ">Figure 10
<p>Comparison of DPPH radical scavenging activity of the peptides at varying concentrations.</p>
Full article ">
45 pages, 11195 KiB  
Review
Exploring Plasmonic Standalone Surface-Enhanced Raman Scattering Nanoprobes for Multifaceted Applications in Biomedical, Food, and Environmental Fields
by Valentina Rojas Martínez, Eunseo Lee and Jeong-Wook Oh
Nanomaterials 2024, 14(22), 1839; https://doi.org/10.3390/nano14221839 - 17 Nov 2024
Viewed by 442
Abstract
Surface-enhanced Raman scattering (SERS) is an innovative spectroscopic technique that amplifies the Raman signals of molecules adsorbed on rough metal surfaces, making it pivotal for single-molecule detection in complex biological and environmental matrices. This review aims to elucidate the design strategies and recent [...] Read more.
Surface-enhanced Raman scattering (SERS) is an innovative spectroscopic technique that amplifies the Raman signals of molecules adsorbed on rough metal surfaces, making it pivotal for single-molecule detection in complex biological and environmental matrices. This review aims to elucidate the design strategies and recent advancements in the application of standalone SERS nanoprobes, with a special focus on quantifiable SERS tags. We conducted a comprehensive analysis of the recent literature, focusing on the development of SERS nanoprobes that employ novel nanostructuring techniques to enhance signal reliability and quantification. Standalone SERS nanoprobes exhibit significant enhancements in sensitivity and specificity due to optimized hot spot generation and improved reporter molecule interactions. Recent innovations include the development of nanogap and core–satellite structures that enhance electromagnetic fields, which are crucial for SERS applications. Standalone SERS nanoprobes, particularly those utilizing indirect detection mechanisms, represent a significant advancement in the field. They hold potential for wide-ranging applications, from disease diagnostics to environmental monitoring, owing to their enhanced sensitivity and ability to operate under complex sample conditions. Full article
(This article belongs to the Special Issue Versatile Plasmonic Nanostructures for Biomedical Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) SERS-based sensors: Direct SERS detection using the Raman signal of target molecules and indirect SERS detection using the Raman signal transformation of Raman reporter molecules (e.g., Raman reporter modified aptamer) or the amplified Raman signal of standalone SERS nanotags. (<b>B</b>) Various nanostructures for efficient standalone SERS nanoprobes.</p>
Full article ">Figure 2
<p>SERS mechanisms: (<b>A</b>) LSPR-induced EM enhancement. (<b>B</b>) CT resonance mechanism of CM enhancement mechanisms at a metal-molecule or semiconductor-molecule interface. The arrows indicate CT transitions (<span class="html-italic">μ</span><sub>CT</sub>), electronic transitions of a molecule (<span class="html-italic">μ</span><sub>mol</sub>), E<sub>F</sub> (Fermi level), HOMO (highest occupied molecular orbital), LUMO (lowest unoccupied molecular orbital), VB (valence band), and CB (conduction band).</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>) Transmission electron microscopy (TEM) images of diverse morphologies of nanoparticles: (<b>A</b>) Au nanosphere (adapted with permission from [<a href="#B68-nanomaterials-14-01839" class="html-bibr">68</a>]; Copyright 2021 American Chemical Society). (<b>B</b>) Au nanotriangles (adapted with permission from [<a href="#B69-nanomaterials-14-01839" class="html-bibr">69</a>]; Copyright 2022 American Chemical Society). (<b>C</b>) Au nanocubes (adapted with permission from [<a href="#B70-nanomaterials-14-01839" class="html-bibr">70</a>]; Copyright 2022 Elsevier). (<b>D</b>) Au nanorods with various aspect ratios. The ratios of Au NRs are 5.9, 6.4, 6.4, 7.5, and 8.5, corresponing to Figures (<b>D</b>-<b>a</b>) through (<b>D</b>-<b>e</b>), respectively. (<b>D</b>-<b>f</b>) UV-vis-NIR spectra of Au NRs shown in (<b>D</b>-<b>a</b>) (black), (<b>D</b>-<b>b</b>) (green), (<b>D</b>-<b>c</b>) (red), (<b>D</b>-<b>d</b>) (blue), and (<b>D</b>-<b>e</b>) (magenta), respectively. The orange curve is the UV-vis-NIR spectrum of Au NRs synthesized with the ratio of 7.3. All scale bars represent 100 nm. (adapted with permission from [<a href="#B63-nanomaterials-14-01839" class="html-bibr">63</a>]; Copyright 2012 American Chemical Society). (<b>E</b>) TEM image of Ag@NO<sub>2</sub> (adapted with permission from [<a href="#B66-nanomaterials-14-01839" class="html-bibr">66</a>]; Copyright 2022 The Royal Society of Chemistry). (<b>F</b>) Au nanostars (adapted with permission from [<a href="#B67-nanomaterials-14-01839" class="html-bibr">67</a>]; Copyright 2012 IOP Publishing).</p>
Full article ">Figure 4
<p>(<b>A</b>) Au dimers with nanogaps bridged by metal-organic molecular cages (MOCs) of different sizes (MOC1, MOC2, and MOC3). (<b>A</b>-<b>a</b>) TEM images, (<b>A</b>-<b>b</b>) HRTEM images, and (<b>A</b>-<b>c</b>) simulated electric field distributions around the dimers. <b>1</b>, <b>2</b>, and <b>3</b> corresponds to MOC1, MOC2, and MOC3, respectively. (adapted with permission from REF [<a href="#B76-nanomaterials-14-01839" class="html-bibr">76</a>]; Copyright 2021 American Chemical Society). (<b>B</b>) DNA origami nanofork-based dimeric structures with various NPs (adapted with permission from REF [<a href="#B79-nanomaterials-14-01839" class="html-bibr">79</a>]; Copyright 2023 American Chemical Society). (<b>C</b>) Dimeric structure with a nanocube and a nanosphere (adapted with permission from REF [<a href="#B80-nanomaterials-14-01839" class="html-bibr">80</a>]; Copyright 2021 Wiley-VCH). (<b>D</b>) Au@Ag nanostar dimer (adapted with permission from REF [<a href="#B81-nanomaterials-14-01839" class="html-bibr">81</a>]; Copyright 2021 American Chemical Society). (<b>E</b>) Detection of endotoxin by SERS chip with dimeric SERS nanotags (adapted with permission from REF [<a href="#B83-nanomaterials-14-01839" class="html-bibr">83</a>]; Copyright 2020 American Chemical Society). (<b>F</b>) Raman imaging of cancer cells with Au dimers (adapted with permission from REF [<a href="#B84-nanomaterials-14-01839" class="html-bibr">84</a>]; Copyright 2017 American Chemical Society). (<b>G</b>) Au dimers, trimers, and comparison of their Raman signals (adapted with permission from REF [<a href="#B87-nanomaterials-14-01839" class="html-bibr">87</a>]; Copyright 2017 Royal Society of Chemistry) (<b>H</b>) DNA origami-based tetramer structure (adapted with permission from REF [<a href="#B86-nanomaterials-14-01839" class="html-bibr">86</a>]; Copyright 2014 American Chemical Society).</p>
Full article ">Figure 5
<p>(<b>A</b>) TEM images of Au@4-MBN@AgNPs with Ag shell thickness of 2.2, 3.6, 6.4, 8.9, 10.1, and 12.2 nm (adapted under the terms of CC-BY License from REF [<a href="#B91-nanomaterials-14-01839" class="html-bibr">91</a>]; Copyright 2024 The Authors, published in Frontiers). (<b>B</b>) Raman intensity of different shell thicknesses of Au@4-MBN@AgNPs at 2221cm<sup>−1</sup> (adapted under the terms of CC-BY License from REF [<a href="#B91-nanomaterials-14-01839" class="html-bibr">91</a>]; Copyright 2024 The Authors, published in Frontiers). (<b>C</b>) HRTEM images of Au@ATP@Ag nanorods obtained at a sub-threshold 4-ATP concentration CATP = 2.0 × 10<sup>−7</sup> M (adapted with permission from REF [<a href="#B92-nanomaterials-14-01839" class="html-bibr">92</a>]; Copyright 2016 Tsinghua University Press and Springer-Verlag GmbH Germany). (<b>D</b>) SERS spectra of the Au@Ag@ATP7 (left) and Au@ATP@Ag7 (right) samples before and after oxidation of the amino groups with hydrogen peroxide. The asterisk represents four additional peaks observed after oxidation, with three peaks at higher wavenumbers corresponding to nitrobenzene (adapted with permission from REF [<a href="#B92-nanomaterials-14-01839" class="html-bibr">92</a>]; Copyright 2016 Tsinghua University Press and Springer-Verlag GmbH Germany). (<b>E</b>) HRTEM images of Au/SiO<sub>2</sub> core–shell nanoparticles, SHINERS: shell-isolated mode and schematic of a SHINERS experiment on living yeast cells (adapted with permission from REF [<a href="#B93-nanomaterials-14-01839" class="html-bibr">93</a>]; Copyright 2010 Springer nature). (<b>F</b>) Schematic representation of H<sub>2</sub>O<sub>2</sub> triggered degradation of MnO<sub>2</sub> coating, TEM image, and evaluation MnO<sub>2</sub> degradation SERS fingerprinting (adapted with permission from REF [<a href="#B94-nanomaterials-14-01839" class="html-bibr">94</a>]; Copyright 2021 American Chemical Society).</p>
Full article ">Figure 6
<p>Synthetic schematic diagram (<b>A</b>) and electric field distribution (<b>B</b>) of SiO<sub>2</sub>@Au-Ag CJS (adapted with permission from REF [<a href="#B97-nanomaterials-14-01839" class="html-bibr">97</a>]; Copyright 2023 American Chemical Society). (<b>C</b>) Schematic diagram of SERS-ELISA platform with CS@SiO<sub>2</sub> core–satellite Au NPs (adapted with permission from REF [<a href="#B100-nanomaterials-14-01839" class="html-bibr">100</a>]; Copyright 2023 Elsevier). UV-vis spectra, TEM images (inset) (<b>D</b>), and SERS spectra (<b>E</b>) of the nanosensor before and after incubation with MMP-2. The characteristic peaks of DTNB (5,5′-dithiobis(2-nitrobenzoic acid)) at 1324 cm<sup>−1</sup> (red dash) and MBN (4-mercaptobenzonitrile) at 1580 cm<sup>−1</sup> (blue range) (adapted with permission from REF [<a href="#B103-nanomaterials-14-01839" class="html-bibr">103</a>]; Copyright 2024 American Chemical Society).</p>
Full article ">Figure 7
<p>(<b>A</b>) Calculated near-field EM field distribution of the Au-NNP and a silica-gapped Au-Au core-gap-shell nanoparticle without a bridge (adapted with permission from [<a href="#B104-nanomaterials-14-01839" class="html-bibr">104</a>]; Copyright 2011 Springer Nature). (<b>B</b>) TEM images of Au-NNP structures after Au shell formation on various DNA-modified Au cores (adapted with permission from [<a href="#B105-nanomaterials-14-01839" class="html-bibr">105</a>]; Copyright 2014 American Chemical Society). (<b>C</b>) Calculated near-field EM field distribution of Au-NNPs with different surface morphologies (adapted with permission from [<a href="#B107-nanomaterials-14-01839" class="html-bibr">107</a>]; Copyright 2016 Wiley-VCH). (<b>D</b>) P-GERTs and S-GERTs (adapted with permission from [<a href="#B110-nanomaterials-14-01839" class="html-bibr">110</a>]; Copyright 2019 Springer Nature). (<b>E</b>) Schematic diagram of high-speed cell Raman imaging and bright-field and Raman images of a single H1299 cell with different parts randomly selected (point 1–3). Scale bars are 10 μm (adapted with permission from [<a href="#B110-nanomaterials-14-01839" class="html-bibr">110</a>]; Copyright 2019 Springer Nature). (<b>F</b>) Progression of structural complexity in nanoframes with increasing chemical steps (adapted with permission from [<a href="#B114-nanomaterials-14-01839" class="html-bibr">114</a>]; Copyright 2023 American Chemical Society). (<b>G</b>) Synthetic scheme and TEM images of AuDGNs (adapted with permission from [<a href="#B115-nanomaterials-14-01839" class="html-bibr">115</a>]; Copyright 2016 Wiley-VCH). (<b>H</b>) OXNCs with different gap sizes and those HAADF-STEM images (i–iii). The scale bars indicate 100 nm (adapted with permission from [<a href="#B113-nanomaterials-14-01839" class="html-bibr">113</a>]; Copyright 2024 American Chemical Society). (<b>I</b>) Structures and sizes of hemin, myoglobin, and hemoglobin (adapted with permission from [<a href="#B113-nanomaterials-14-01839" class="html-bibr">113</a>]; Copyright 2024 American Chemical Society). (<b>J</b>) SERS spectra of hemin (green line) mixed with the OXNC with 2.6 nm gaps, myoglobin (blue line) with the OXNC with 5.6 nm gaps, and hemoglobin (orange line) mixed with the OXNC with 5.6 nm gaps (adapted with permission from [<a href="#B113-nanomaterials-14-01839" class="html-bibr">113</a>]; Copyright 2024 American Chemical Society).</p>
Full article ">Figure 8
<p>(<b>A</b>) ERA-SERS-LF strip (left) and phototgraphs (right, inset) and the calibration curve (Right) of SiO<sub>2</sub>@Au-based ERA-LF-SERS strips when testing the IAV DNA (adapted with permission from [<a href="#B120-nanomaterials-14-01839" class="html-bibr">120</a>]; Copyright 2023 American Chemical Society). (<b>B</b>) Clinical serum sample tests by Ag@Au NP-based dual-mode LFIA (adapted with permission from [<a href="#B121-nanomaterials-14-01839" class="html-bibr">121</a>]; Copyright 2022 American Chemical Society). (<b>C</b>) Schematic representation of assay. Total RNA is first isolated from samples before target RNA biomarkers are simultaneously amplified using isothermal reverse transcription-recombinase polymerase amplification. During amplification, amplicons are tagged with biotin molecules and target-specific overhang hybridization sequences. The different biomarker-specific amplicons are then labeled with respective SERS nanotags through complementary sequence hybridization and magnetically purified. Finally, the amplicons are detected by SERS concurrently, and quantitative analysis of biomarker level is derived from the spectral peak of each unique SERS nanotag. The Raman signals correspond to characteristic peaks from the five different dyes of the SERS nanotags, respectively. (Adapted with permission from [<a href="#B126-nanomaterials-14-01839" class="html-bibr">126</a>]; Copyright 2016 Wiley-VCH).</p>
Full article ">Figure 9
<p>(<b>A</b>) Multiplexed biomarker detection using ER, PR, and HER2 IgGs conjugated SERS nanotags (adapted with permission from [<a href="#B130-nanomaterials-14-01839" class="html-bibr">130</a>]; Copyright 2023 Elsevier). (<b>B</b>) Colors of Au<sub>13</sub>NPs, ASNPs, AS@mSiO<sub>2</sub> NPs, and pAS@AuNCs suspended in nanopure water with SPR peaks at 518, 700, 734, and 806 nm, respectively (adapted with permission from [<a href="#B131-nanomaterials-14-01839" class="html-bibr">131</a>]; Copyright 2023 Wiley-VCH). (<b>C</b>) Schematic illustration application of the multilayered mesoporous Au nanoarchitecture (RGD/DOX-pAS@AuNC) labeled with Raman reporter (MBA) via Au–thiol covalent bond for surface-enhanced Raman scattering (SERS) imaging-guided synergistic therapy toward cancer. (adapted with permission from [<a href="#B131-nanomaterials-14-01839" class="html-bibr">131</a>]; Copyright 2023 Wiley-VCH). (<b>D</b>) Schematic illustration showing that AuDAg<sub>2</sub>S nanoprobes equipped with SERS/NIR-II optical imaging could multidimensional tumor images from living subjects, pathology to the single-cell and further guided NIR-II deeper photothermal therapy (adapted with permission from [<a href="#B132-nanomaterials-14-01839" class="html-bibr">132</a>]; Copyright 2022 Wiley-VCH). (<b>E</b>) Fabrication of Oligonucleotide Modified Bioorthogonal SERS Nanotags (adapted with permission from [<a href="#B133-nanomaterials-14-01839" class="html-bibr">133</a>]; Copyright 2020 American Chemical Society). (<b>F</b>) Bioorthogonal SERS nanotags as a precision theranostic platform for cancer detection and photothermal therapy in mice after intravenous injection (adapted with permission from [<a href="#B133-nanomaterials-14-01839" class="html-bibr">133</a>]; Copyright 2020 American Chemical Society). (<b>G</b>) Photographic image of a BALB/c mouse with blank and <span class="html-italic">S. aureus</span> infected wounds after applying ACPA and SERS images at 2086 cm<sup>−1</sup> of <span class="html-italic">S. aureus</span> (right) and blank (left) infected wounds at different time points (left). Corresponding average SERS intensities of ACPA on wounds. *** <span class="html-italic">p</span> &lt; 0.001 (right) (adapted with permission from [<a href="#B134-nanomaterials-14-01839" class="html-bibr">134</a>]; Copyright 2023 American Chemical Society).</p>
Full article ">Figure 10
<p>(<b>A</b>) SERS spectral responses obtained from the reaction of the developed SERS aptasensor with various concentrations of pathogens (adapted with permission from [<a href="#B135-nanomaterials-14-01839" class="html-bibr">135</a>]; Copyright 2020 Elsevier). (<b>B</b>) Photographs of the LFIA strip with histamine (Hist), parvalbumin (Parv), and protein-G (PG) immobilized in the test (T) and control (<b>C</b>) lines, as indicated. SERS intensity mappings acquired at 1616 and 1646 cm<sup>−1</sup>, which are characteristic peaks of αHist-MGITC SERS or αParvRBITC SERS tags, respectively. (<b>C</b>) Average SERS spectra acquired from the different concentrations of histamine. (<b>D</b>) Average SERS spectra acquired from the different concentrations of Parvalbumin ((<a href="#nanomaterials-14-01839-f007" class="html-fig">Figure 7</a>B–D) adapted with permission from [<a href="#B136-nanomaterials-14-01839" class="html-bibr">136</a>]; Copyright 2024 American Chemical Society). (<b>E</b>) Photographs of competitive LFIA (CLFIA) strips at different concentrations of AFB<sub>1</sub>. The black arrow marks the T-line, indicating the visible LOD (i.e., 0.2 ng/mL) as determined by 12 independent users using only the naked eye. (<b>F</b>) Photographs (left) and SEM images (right) of the CLFIA strip membrane at the AFB<sub>1</sub> concentrations of (<b>i</b>) 0 ng/mL, (<b>ii</b>) 0.05 ng/mL, and (<b>iii</b>) 0.2 ng/mL. The blue arrows annotate Au-Ag alloy NPs-incorporated silica spheres captured in the T-line, with their number gradually decreasing as AFB<sub>1</sub> concentration increases. No nanoparticles are observed in (<b>iii</b>). ((<b>E</b>,<b>F</b>) adapted with permission from [<a href="#B139-nanomaterials-14-01839" class="html-bibr">139</a>]; Copyright 2023 American Chemical Society). (<b>G</b>) Scheme of the SERS microarray immunoassay for multiple mycotoxins (adapted with permission from [<a href="#B140-nanomaterials-14-01839" class="html-bibr">140</a>]; Copyright 2024 American Chemical Society).</p>
Full article ">Figure 11
<p>(<b>A</b>) Aptamer-based turn-off dual SERS sensor with AuNF-Au@tag@Ag@Au NP core-satellite assembly platform for MC-LR and MC-RR (L: leucine, R: arginine). (<b>B</b>) Optical brightfield image of <span class="html-italic">M. aeruginosa</span> UTEX LB 2385 cells. (<b>C</b>). MC-LR levels produced by <span class="html-italic">M. aeruginosa</span> UTEX LB 2385 (curve a) and <span class="html-italic">C. reinhardti</span> (curve b) over 7 consecutive days, as determined by the aptasensor ((<b>A</b>–<b>C</b>) adapted with permission from [<a href="#B141-nanomaterials-14-01839" class="html-bibr">141</a>]; Copyright 2021 American Chemical Society). (<b>D</b>) Schematic diagram of the optical setup of the SPR-SERS microscope and detecting strategy for Pb<sup>2+</sup> and Hg<sup>2+</sup> using single-particle Raman imaging (adapted with permission from [<a href="#B142-nanomaterials-14-01839" class="html-bibr">142</a>]; Copyright 2023 American Chemical Society). (<b>E</b>) SERS-based AMP immunoassay with magnetic separation (adapted with permission from [<a href="#B143-nanomaterials-14-01839" class="html-bibr">143</a>]; Copyright 2022 Royal Society of Chemistry). (<b>F</b>) Detection of series BPA actual samples using the SERS ICA (ICA: immunochromatographic assay) strips (adapted with permission from [<a href="#B144-nanomaterials-14-01839" class="html-bibr">144</a>]; Copyright 2022 Elsevier). (<b>G</b>) An image of the detected organs of a bivalve <span class="html-italic">Ruditapes philippinarum</span>, Au NS@polystyrene (PS) core@shell structures with Cy7 dyes, and typical SERS spectra measured from the organs of the clams exposed to SERS@PS for 24 h. (adapted with permission from [<a href="#B145-nanomaterials-14-01839" class="html-bibr">145</a>]; Copyright 2022 Royal Society of Chemistry).</p>
Full article ">Figure 12
<p>(<b>A</b>) Schematic illustration of the synthesis process of polydopamine@gold (PDA@Au) nanowaxberry and its SERS detection. (I) Deposition of Au seeds onto the surface of the PDA sphere, (II) the iodide ions assisted the growth of Au nanoshell on the PDA sphere, and (III) SERS detection of pesticides, pollutants, and explosives using nanowaxberry as a substrate (adapted with permission from [<a href="#B147-nanomaterials-14-01839" class="html-bibr">147</a>]; Copyright 2018 American Chemical Society). (<b>B</b>) Schematic of the SERS nanosensor for •OH detection and mechanism and detection of H<sub>2</sub>O<sub>2</sub> and •OH generation in water microdroplets (adapted with permission from [<a href="#B148-nanomaterials-14-01839" class="html-bibr">148</a>]; Copyright 2024 American Chemical Society).</p>
Full article ">
13 pages, 3968 KiB  
Article
Insight into the Local Surface Plasmon Resonance Effect of Pt-SnS2 Nanosheets in Tetracycline Photodegradation
by Mao Feng, Tianhao Zhou, Jiaxin Li, Mengqing Cao, Jing Cheng, Danyang Li, Jian Qi and Feifei You
Molecules 2024, 29(22), 5423; https://doi.org/10.3390/molecules29225423 - 17 Nov 2024
Viewed by 409
Abstract
Constructing highly efficient catalysts for the degradation of organic pollutants driven by solar light in aquatic environments is a promising and green strategy. In this study, a novel hexagonal sheet-like Pt/SnS2 heterojunction photocatalyst is successfully designed and fabricated using a hydrothermal method [...] Read more.
Constructing highly efficient catalysts for the degradation of organic pollutants driven by solar light in aquatic environments is a promising and green strategy. In this study, a novel hexagonal sheet-like Pt/SnS2 heterojunction photocatalyst is successfully designed and fabricated using a hydrothermal method and photodeposition process for photocatalytic tetracycline (TC) degradation. The optimal Pt/SnS2 hybrid behaves with excellent photocatalytic performance, with a degradation efficiency of 91.27% after 120 min, a reaction rate constant of 0.0187 min−1, and durability, which can be attributed to (i) the formation of a metal/semiconductor interface field caused by loading Pt nanoparticles (NPs) on the surface of SnS2, facilitating the separation of photo-induced charge carriers; (ii) the local surface plasmon resonance (LSPR) effect of Pt NPs, extending the light absorption range; and (iii) the sheet-like structure of SnS2, which can shorten the transmission distance of charge carriers, thereby allowing more electrons (e) and holes (h+) to transfer to the surface of the catalyst. This work provides new insights with the utilization of sheet-like structured materials for highly active photocatalytic TC degradation in wastewater treatment and environmental remediation. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the synthesis process of sheet-like-structured SnS<sub>2</sub> and Pt/SnS<sub>2</sub> hybrids; TEM images of (<b>b</b>) SnO<sub>2</sub> hollow sphere, (<b>c</b>) SnS<sub>2</sub> and (<b>d</b>) SnS<sub>2</sub>-2.0Pt with inset of particle size distribution of Pt NPs; HRTEM image of (<b>e</b>) SnS<sub>2</sub>-2.0Pt; (<b>f</b>) HAADF−STEM and (<b>g</b>–<b>i</b>) elemental distribution images of SnS<sub>2</sub>−2.0Pt; and (<b>j</b>) XRD patterns of SnS<sub>2</sub> and SnS<sub>2</sub>−2.0Pt with standard diffraction peaks of SnS<sub>2</sub> and Pt (vertical lines).</p>
Full article ">Figure 2
<p>XPS spectra of SnS<sub>2</sub> and SnS<sub>2</sub>−2.0Pt: (<b>a</b>) survey; high-resolution XPS spectra of (<b>b</b>) Sn 3d, (<b>c</b>) S 2p, and (<b>d</b>) Pt 4f.</p>
Full article ">Figure 3
<p>(<b>a</b>) Photocatalytic activities of all as-prepared samples during the degradation of TC, (<b>b</b>) kinetic curves, (<b>c</b>) the reaction rate constant, and (<b>d</b>) durability tests of SnS<sub>2</sub>−2.0Pt.</p>
Full article ">Figure 4
<p>Photoelectronic characterizations of SnS<sub>2</sub> and SnS<sub>2</sub>−2.0Pt: (<b>a</b>) steady-state PL spectra, (<b>b</b>) time-resolved PL spectra, (<b>c</b>) EIS Nyquist plots and the fitting circuit diagram (inset), and (<b>d</b>) photocurrent density−time curves.</p>
Full article ">Figure 5
<p>(<b>a</b>) Photocatalytic activities of the degradation of TC and (<b>b</b>) TC removal efficiency with the SnS<sub>2</sub>−2.0Pt catalyst in the presence of various scavengers.</p>
Full article ">Figure 6
<p>(a) UV−Vis spectra of all samples, (<b>b</b>) Tuac curves of SnS<sub>2</sub> and SnS<sub>2</sub>−2.0Pt, (<b>c</b>) Mott−Schottky plots of SnS<sub>2</sub> and (<b>d</b>) the main proposed photocatalytic TC degradation mechanism diagram of the sheet-like SnS<sub>2</sub>−2.0Pt heterogeneous catalyst.</p>
Full article ">
16 pages, 4312 KiB  
Article
Peptide-Functionalized Gold Nanoparticles as Organocatalysts for Asymmetric Aldol Reactions
by Thabo Peme, Dean Brady, Ndivhuwo P. Shumbula, Khanani Machumele, Nosipho Moloto, Taryn Adams and Maya M. Makatini
Catalysts 2024, 14(11), 826; https://doi.org/10.3390/catal14110826 - 16 Nov 2024
Viewed by 431
Abstract
The use of high catalyst loading is required for most of the organocatalyzed asymmetric aldol reactions in organic synthesis, and this often presents challenges during purification and difficulties in catalyst recovery from the reaction mixture. The immobilization of the catalyst onto gold nanoparticles [...] Read more.
The use of high catalyst loading is required for most of the organocatalyzed asymmetric aldol reactions in organic synthesis, and this often presents challenges during purification and difficulties in catalyst recovery from the reaction mixture. The immobilization of the catalyst onto gold nanoparticles (AuNPs) can change the structural conformations of the catalyst, thereby improving its catalytic activity and reusability. Herein we report on the synthesis of aldolase mimetic peptide coupled to gold nanoparticles (AuNPs) as efficient organocatalysts for asymmetric aldol reaction. AuNPs were synthesized using the Turkevich method. The conjugation of the peptide to AuNPs was characterized using surface plasmon resonance (SPR), Raman and X-ray photoelectron spectroscopy, and transmission electron microscopy (TEM) was used for particle size determination. The produced nanoparticles, whose sizes depended on the reduction method, were quasi-spherical with a relatively narrow size distribution. The peptide–AuNP conjugates were evaluated for aldol reaction catalytic activity between carbonyls p-nitrobenzaldehyde and cyclohexanone. The products were obtained with good yields (up to 85%) and enantioselectivity (up to 94%). The influence of organic solvents, pH and buffer solutions was also investigated. The results showed that the buffer solutions regulated the colloidal stability of AuNPs, resulting in a significant enhancement in the catalytic rate of the peptide–AuNP conjugate. Full article
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the peptide.</p>
Full article ">Figure 2
<p>UV–vis absorption spectra showing the LSPR band for citrate-capped AuNPs (black line) and peptide–AuNPs (red line).</p>
Full article ">Figure 3
<p>TEM images of citrate–AuNPs and peptide–AuNPs: (<b>a</b>) AuNP-20, (<b>b</b>) AuNP-30, (<b>c</b>) AuNP-30, (<b>d</b>) AuNP-20P, (<b>e</b>) AuNP-30P and (<b>f</b>) AuNP-50P.</p>
Full article ">Figure 4
<p>Particle distribution histogram of citrate-capped AuNPs and peptide–AuNPs.</p>
Full article ">Figure 5
<p>X-ray photoelectron survey and high-resolution (O1s, Au4f and valence) spectra of citrate-capped AuNPs (black line) and peptide–AuNPs (red line).</p>
Full article ">Figure 6
<p>(<b>A</b>) Circular dichroism (CD) spectra in water for unbound peptide (TP_ADLys-W) and bounded peptide using 16 nm (Au20P-W) and 12 nm (Au50P-W) nanoparticles; (<b>B</b>) circular dichroism (CD) spectra in buffer for unbound peptide (TP_ADLys-Bf) and bounded peptide using 16 nm (Au20P-Bf) and 12 nm (Au50P-Bf) nanoparticles, respectively.</p>
Full article ">Figure 6 Cont.
<p>(<b>A</b>) Circular dichroism (CD) spectra in water for unbound peptide (TP_ADLys-W) and bounded peptide using 16 nm (Au20P-W) and 12 nm (Au50P-W) nanoparticles; (<b>B</b>) circular dichroism (CD) spectra in buffer for unbound peptide (TP_ADLys-Bf) and bounded peptide using 16 nm (Au20P-Bf) and 12 nm (Au50P-Bf) nanoparticles, respectively.</p>
Full article ">Scheme 1
<p>Aldol reactions catalyzed by FBPA peptide mimics [<a href="#B35-catalysts-14-00826" class="html-bibr">35</a>].</p>
Full article ">Scheme 2
<p>Schematic representation of peptide–AuNPs synthesized from citrate-capped gold nanoparticles.</p>
Full article ">
11 pages, 3194 KiB  
Article
Theoretical Design of Smart Bionic Skins with Self-Adaptive Temperature Regulation
by Yubo Wang, Yungui Ma and Rui Chen
Materials 2024, 17(22), 5580; https://doi.org/10.3390/ma17225580 - 15 Nov 2024
Viewed by 312
Abstract
Thermal management presents a significant challenge in electric design, particularly in densely packed electronic systems. This study proposes a theoretical model for radiative bionic skin that emulates human skin, enabling the self-adaptive modulation of the thermal exhaustion rate to maintain homeostasis for objects [...] Read more.
Thermal management presents a significant challenge in electric design, particularly in densely packed electronic systems. This study proposes a theoretical model for radiative bionic skin that emulates human skin, enabling the self-adaptive modulation of the thermal exhaustion rate to maintain homeostasis for objects covered by the skin in fluctuating thermal environments. The proposed artificial skin consists of phase change material (VO2) nanoparticles embedded in a low-loss matrix situated on a metallic substrate with a minimal thickness of several micrometers. The findings from our theoretical analyses indicate that substantial alterations in thermal radiation power around the phase transition temperature of 340 K enable a silicone substrate to sustain a relatively stable temperature, with variations confined to ±6 K, despite external heat fluxes ranging from 150 to 450 W/m2. Furthermore, to improve the spectral resemblance to natural skin, a plasmonic surface composed of self-assembled silver nanocubes is incorporated, allowing for modifications to the visible light properties of the bionic skin while maintaining its infrared characteristics. This theoretical investigation offers a cost-effective and conformal approach to the design of ultra-compact, fully passive, and versatile thermal management solutions for robotic systems and related technologies. Full article
Show Figures

Figure 1

Figure 1
<p>An illustration of self-adaptive temperature regulation. The red and blue arm regions covered by the smart skin coat represent the situations with high and low external heat loads, respectively. The red arrows denote the thermal radiation process, with the arrow thickness representing radiation strength. The radiation is automatically strengthened (weakened) with a high (low) heat load.</p>
Full article ">Figure 2
<p>Schematics of the designed structures. (<b>a</b>) The designed structure composed of a hybrid film (mixture of VO<sub>2</sub> nano-inclusions in the PE matrix) and a gold back plate. (<b>b</b>) The reference structure is composed of a pure VO<sub>2</sub> film and a gold back plate. The substrate in both samples is made of silicone. The thicknesses for each layer are <span class="html-italic">t</span><sub>1</sub> = 500 μm, <span class="html-italic">t</span><sub>2</sub> = 0.2 μm, <span class="html-italic">t</span><sub>3</sub> = 1.13 μm, and <span class="html-italic">t</span><sub>4</sub> = 0.41 μm, respectively.</p>
Full article ">Figure 3
<p>Effective index parameters for (<b>a</b>) hybrid film and (<b>b</b>) pure VO<sub>2</sub> film. The solid lines are the real part of the refractive index and the dashed lines denote the imaginary part of the index. The red and black colors represent the insulating and metallic phase states of VO<sub>2</sub>, respectively.</p>
Full article ">Figure 4
<p>Infrared spectra emissivity of the samples at different wave emission angles. (<b>a</b>,<b>b</b>) The emissivity of the hybrid film structure with insulating and metallic VO<sub>2</sub>, respectively. (<b>c</b>,<b>d</b>) The emissivity of the reference VO<sub>2</sub> film structure with insulating and metallic VO<sub>2</sub>, respectively. Colors represent the value of emissivity.</p>
Full article ">Figure 5
<p>Thermal radiation power of the samples at different temperatures. (<b>a</b>) The radiation response of the optimized structures for both hybrid and VO<sub>2</sub> films. (<b>b</b>) The thermal radiation properties of the hybrid film structure at different combinations of the filling factor <span class="html-italic">f</span> and the film thickness <span class="html-italic">t</span><sub>3</sub>. Hysteresis behavior is exhibited by solid and dashed lines, respectively, representing the heating and cooling process. The blackbody radiation (red solid) is also given for comparison.</p>
Full article ">Figure 6
<p>Temperature fluctuation at different input powers for the samples. The input power is 150–450 W/m<sup>2</sup> for (<b>a</b>), 0–200 W/m<sup>2</sup> for (<b>b</b>), and 300–700 W/m<sup>2</sup> for (<b>c</b>), and the corresponding transient temperature responses for the hybrid sample (blue solid) and the VO<sub>2</sub> reference (red solid) are given in (<b>d</b>), (<b>e</b>), and (<b>f</b>), respectively. In (<b>d</b>,<b>f</b>), the temperature fluctuation range at the homeostasis is also added.</p>
Full article ">Figure 7
<p>Visible and infrared spectra properties. (<b>a</b>) Reflectivity in the visible light spectra at various points of the silver nanocube. The inset gives one unit of the silver nanocube, comprising the top plasmonic surface. (<b>b</b>) Emissivity properties of the infrared spectrum. The normal light incidence and emissions are calculated here.</p>
Full article ">
17 pages, 6624 KiB  
Article
Laser-Induced Silver Nanowires/Polymer Composites for Flexible Electronics and Electromagnetic Compatibility Application
by Il’ya Bril’, Anton Voronin, Yuri Fadeev, Alexander Pavlikov, Ilya Govorun, Ivan Podshivalov, Bogdan Parshin, Mstislav Makeev, Pavel Mikhalev, Kseniya Afanasova, Mikhail Simunin and Stanislav Khartov
Polymers 2024, 16(22), 3174; https://doi.org/10.3390/polym16223174 - 14 Nov 2024
Viewed by 474
Abstract
Nowadays, the Internet of Things (IOT), electronics, and neural interfaces are becoming an integral part of our life. These technologies place unprecedentedly high demands on materials in terms of their mechanical and electrical properties. There are several strategies for forming conductive layers in [...] Read more.
Nowadays, the Internet of Things (IOT), electronics, and neural interfaces are becoming an integral part of our life. These technologies place unprecedentedly high demands on materials in terms of their mechanical and electrical properties. There are several strategies for forming conductive layers in such composites, e.g., volume blending to achieve a percolation threshold, inkjet printing, lithography, and laser processing. The latter is a low-cost, environmentally friendly, scalable way to produce composites. In our work, we synthesized AgNW and characterized them using Ultraviolet-visible spectroscopy (UV-vis), Transmission electron microscopy (TEM), and Selective area electron diffraction (SAED). We found that our AgNW absorbed in the UV-vis range of 345 to 410 nm. This is due to the plasmon resonance phenomenon of AgNW. Then, we applied the dispersion of AgNW on the surface of the polymer substrate, dried them and we got the films of AgNW.. We irradiated these films with a 432 nm laser. As a result of the treatment, we observed two processes. The first one was the sintering and partial melting of nanowires under the influence of laser radiation, as a consequence of which, the sheet resistance dropped more than twice. The second was the melting of the polymer at the interface and the subsequent integration of AgNW into the substrate. This allowed us to improve the adhesion from 0–1 B to 5 B, and to obtain a composite capable of bending, with radius of 0.5 mm. We also evaluated the shielding efficiency of the obtained composites. The shielding efficiency for 500–600 nm thick porous film samples were 40 dB, and for 3.1–4.1 µm porous films the shielding efficiency was about 85–90 dB in a frequency range of 0.01–40 GHz. The data obtained by us are the basis for producing flexible electronic components based on AgNW/PET composite for various applications using laser processing methods. Full article
(This article belongs to the Special Issue Multifunctional Polymer Composite Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>AgNW characterization. (<b>a</b>)—TEM Image, (<b>b</b>)—SAED, (<b>c</b>)—AgNW diameter, (<b>d</b>)—AgNW length, (<b>e</b>)—UV-vis absorption spectra.</p>
Full article ">Figure 2
<p>Samples preparation and characterization. (<b>a</b>)—Samples obtaining scheme, (<b>b</b>)—effect of power density on electrical parameters and morphology, (<b>c</b>)—SEM images at different power densities, (<b>d</b>)—XRD of original film, (<b>e</b>)—XRD of laser-processed film.</p>
Full article ">Figure 3
<p>Thickness effect study. (<b>a</b>)—Dependence of thickness from concentration of AgNW basic dispersion, (<b>b</b>)—SEM image of x-section of 62.5 μL/cm2 BSE (top image), (<b>c</b>)—EDX spectra.</p>
Full article ">Figure 4
<p>Electrical properties. (<b>a</b>)—Four-point probe measurements illustration, (<b>b</b>)—sheet resistance/thickness dependence and resistivity of AgNW/PET composite, (<b>c</b>)—sheet resistance/thickness comparison [<a href="#B35-polymers-16-03174" class="html-bibr">35</a>,<a href="#B36-polymers-16-03174" class="html-bibr">36</a>,<a href="#B37-polymers-16-03174" class="html-bibr">37</a>,<a href="#B38-polymers-16-03174" class="html-bibr">38</a>,<a href="#B39-polymers-16-03174" class="html-bibr">39</a>,<a href="#B40-polymers-16-03174" class="html-bibr">40</a>,<a href="#B41-polymers-16-03174" class="html-bibr">41</a>,<a href="#B42-polymers-16-03174" class="html-bibr">42</a>].</p>
Full article ">Figure 5
<p>Electromagnetic compatibility. (<b>a</b>)—Waveguide method illustration, (<b>b</b>)—S<sub>21</sub> parameter at 0.01–7 GHz range, (<b>c</b>)—S<sub>21</sub> parameter at 17–26.5 GHz range, (<b>d</b>)—S<sub>21</sub> parameter at 26.5–40 GHz range.</p>
Full article ">Figure 6
<p>Reflection, absorption, transmission (RAT) diagrams. (<b>a</b>)—RAT at 0.01–7 GHz, (<b>b</b>)—RAT at 17–26.5 GHz, (<b>c</b>)—RAT at 26.5–40 GHz, (<b>d</b>)—AgNW/PET composite EMI shielding illustration.</p>
Full article ">Figure 7
<p>AgNW/PET composite comparisons with alternative materials. (<b>a</b>)—SE comparison [<a href="#B45-polymers-16-03174" class="html-bibr">45</a>,<a href="#B46-polymers-16-03174" class="html-bibr">46</a>,<a href="#B47-polymers-16-03174" class="html-bibr">47</a>,<a href="#B48-polymers-16-03174" class="html-bibr">48</a>,<a href="#B49-polymers-16-03174" class="html-bibr">49</a>,<a href="#B50-polymers-16-03174" class="html-bibr">50</a>,<a href="#B51-polymers-16-03174" class="html-bibr">51</a>,<a href="#B52-polymers-16-03174" class="html-bibr">52</a>,<a href="#B53-polymers-16-03174" class="html-bibr">53</a>,<a href="#B54-polymers-16-03174" class="html-bibr">54</a>,<a href="#B55-polymers-16-03174" class="html-bibr">55</a>,<a href="#B56-polymers-16-03174" class="html-bibr">56</a>,<a href="#B57-polymers-16-03174" class="html-bibr">57</a>,<a href="#B58-polymers-16-03174" class="html-bibr">58</a>,<a href="#B59-polymers-16-03174" class="html-bibr">59</a>,<a href="#B60-polymers-16-03174" class="html-bibr">60</a>,<a href="#B61-polymers-16-03174" class="html-bibr">61</a>], (<b>b</b>)—SSE<sub>T</sub> comparison [<a href="#B61-polymers-16-03174" class="html-bibr">61</a>,<a href="#B62-polymers-16-03174" class="html-bibr">62</a>,<a href="#B63-polymers-16-03174" class="html-bibr">63</a>,<a href="#B64-polymers-16-03174" class="html-bibr">64</a>,<a href="#B65-polymers-16-03174" class="html-bibr">65</a>,<a href="#B66-polymers-16-03174" class="html-bibr">66</a>,<a href="#B67-polymers-16-03174" class="html-bibr">67</a>,<a href="#B68-polymers-16-03174" class="html-bibr">68</a>].</p>
Full article ">Figure 8
<p>Mechanical properties. (<b>a</b>)—Optical image of untreated film before and after the ASTM D3359 test; (<b>b</b>)—optical image of processed film with C<sub>AgNW</sub> = 12.5 μL/cm<sup>2</sup> before and after the ASTM D3359 test; (<b>c</b>)—optical image of processed film with C<sub>AgNW</sub> = 75 μL/cm<sup>2</sup> before and after the ASTMD D3359 test; (<b>d</b>)—change in resistance after a tape test; (<b>e</b>)—changing of resistance at different bending radius; (<b>f</b>)—thermoforming demonstration.</p>
Full article ">
15 pages, 3259 KiB  
Article
Structure and Activity of β-Oligosaccharides Obtained from Lentinus edodes (Shiitake)
by Wei Jia, Wenhan Wang, Yanzhen Yu, Huimin Wang, Hongtao Zhang, Peng Liu, Meiyan Zhang, Qiaozhen Li, Henan Zhang, Huaxiang Li and Jingsong Zhang
Separations 2024, 11(11), 326; https://doi.org/10.3390/separations11110326 - 14 Nov 2024
Viewed by 325
Abstract
The structure and characteristics of LEOPs, β-oligosaccharides from the fruiting body of Lentinus edodes obtained via acid degradation and gel permeation chromatography, were investigated. We performed high-performance liquid chromatography, infrared spectroscopy, methylation analysis, nuclear magnetic resonance, and correlated activity experiments, including antioxidant, immunomodulatory, [...] Read more.
The structure and characteristics of LEOPs, β-oligosaccharides from the fruiting body of Lentinus edodes obtained via acid degradation and gel permeation chromatography, were investigated. We performed high-performance liquid chromatography, infrared spectroscopy, methylation analysis, nuclear magnetic resonance, and correlated activity experiments, including antioxidant, immunomodulatory, and liver injury protection to gain insights. LEOPs comprised an oligosaccharide (Mw 2445 Da) based on six β-1, 3-D-glucose residues as the main chain and six β-1, 6-D-glucose residues as the side chain. Surface plasmon resonance analysis indicated that LEOPs directly bound to dectin-1, which facilitated their immunoenhancing activity via downstream NF-κB activation. The results implied that LEOPs may be the active unit of the shiitake β-glucan. The determination of LEOPs structure was performed to reveal the anti-tumor effect and immune-regulatory function of shiitake β-glucan on a molecular level to provide a basis. Full article
(This article belongs to the Special Issue Research Progress for Isolation of Plant Active Compounds)
Show Figures

Figure 1

Figure 1
<p>High-performance liquid chromatography (HPLC) of LEOPs.</p>
Full article ">Figure 2
<p>Infrared spectra of LEOPs.</p>
Full article ">Figure 3
<p>GC-MS pattern of alditol acetates from the methylation product of LEOPs.</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectra of LEOPs dissolved in DMSO-<span class="html-italic">d6</span> at 25 °C. f1 represents the frequency offset of the observation channel (first channel). The residues A, B, C, and D are arranged from low to high fields.</p>
Full article ">Figure 5
<p><sup>13</sup>C NMR spectra of LEOPs dissolved in DMSO-<span class="html-italic">d6</span> at 25 °C. f1 represents the frequency offset of the observation channel (first channel). The anomeric carbon is labeled based on anomeric hydrogen as A, B, C, and D.</p>
Full article ">Figure 6
<p>Structure of LEOPs.</p>
Full article ">Figure 7
<p>Antioxidant activity of LEOPs based on scavenging DPPH· free radicals. Vitamin C is the positive control. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Stimulatory effect of LEOPs on RAW264.7 phagocytic function. LPS is the positive control. ** and **** stand for significant differences at <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001 levels, respectively, compared with the negative group.</p>
Full article ">Figure 9
<p>Immunostimulatory effect of LEOPs is mediated via NF-κB activation and the dectin-1 receptor. Scleroglucan is the positive control. Different lowercase letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 10
<p>SPR analysis showing binding affinities between LEOPs and dectin-1 receptors. (<b>a</b>) 1:1 steady-state affinity fitting curve for dectin-1 receptor; the determined K<sub>D</sub> value was 340.2 nM; (<b>b</b>) SPR sensograms of LEOPs binding to dectin-1 receptors at different concentrations. The dotted line represents KD value.</p>
Full article ">Figure 11
<p>Repairing effect of LEOPs on alcoholic liver injury (<b>a</b>) and H<sub>2</sub>O<sub>2</sub> liver injury (<b>b</b>). * and ** stand for significant differences at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 levels, respectively, compared with the negative control group. All values are presented as means ± SEM (n = 3). Significantly different (### <span class="html-italic">p</span> &lt; 0.001) versus the CK group, significantly different (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01) versus the H<sub>2</sub>O<sub>2</sub> or EtOH group.</p>
Full article ">
Back to TopTop