[go: up one dir, main page]

Previous Issue
Volume 14, November-1
 
 

Nanomaterials, Volume 14, Issue 22 (November-2 2024) – 78 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
18 pages, 1855 KiB  
Article
Free-Standing Carbon Nanofiber Films with Supported Cobalt Phosphide Nanoparticles as Cathodes for Hydrogen Evolution Reaction in a Microbial Electrolysis Cell
by Gerard Pérez-Pi, Jorge Luque-Rueda, Pau Bosch-Jimenez, Eduard Borràs Camps and Sandra Martínez-Crespiera
Nanomaterials 2024, 14(22), 1849; https://doi.org/10.3390/nano14221849 (registering DOI) - 19 Nov 2024
Abstract
High-performance and cost-efficient electrocatalysts and electrodes are needed to improve the hydrogen evolution reaction (HER) for the hydrogen (H2) generation in electrolysers, including microbial electrolysis cells (MECs). In this study, free-standing carbon nanofiber (CNF) films with supported cobalt phosphide nanoparticles have [...] Read more.
High-performance and cost-efficient electrocatalysts and electrodes are needed to improve the hydrogen evolution reaction (HER) for the hydrogen (H2) generation in electrolysers, including microbial electrolysis cells (MECs). In this study, free-standing carbon nanofiber (CNF) films with supported cobalt phosphide nanoparticles have been prepared by means of an up-scalable electrospinning process followed by a thermal treatment under controlled conditions. The produced cobalt phosphide-supported CNF films show to be nanoporous (pore volume up to 0.33 cm3 g−1) with a high surface area (up to 502 m2 g−1) and with a suitable catalyst mass loading (up to 0.49 mg cm−2). Values of overpotential less than 140 mV at 10 mA cm−2 have been reached for the HER in alkaline media (1 M KOH), which demonstrates a high activity. The high electrical conductivity together with the mechanical stability of the free-standing CNF films allowed their direct use as cathodes in a MEC reactor, resulting in an exceptionally low voltage operation (0.75 V) with a current density demand of 5.4 A m−2. This enabled the production of H2 with an energy consumption below 30 kWh kg−1 H2, which is highly efficient. Full article
(This article belongs to the Special Issue Hydrogen Production and Evolution Based on Nanocatalysts)
25 pages, 10477 KiB  
Article
Portable Homemade Magnetic Hyperthermia Apparatus: Preliminary Results
by Teresa Castelo-Grande, Paulo A. Augusto, Lobinho Gomes, Eduardo Calvo and Domingos Barbosa
Nanomaterials 2024, 14(22), 1848; https://doi.org/10.3390/nano14221848 (registering DOI) - 19 Nov 2024
Abstract
This study aims to describe and evaluate the performance of a new device for magnetic hyperthermia that can produce an alternating magnetic field with adjustable frequency without the need to change capacitors from the resonant bank, as required by other commercial devices. This [...] Read more.
This study aims to describe and evaluate the performance of a new device for magnetic hyperthermia that can produce an alternating magnetic field with adjustable frequency without the need to change capacitors from the resonant bank, as required by other commercial devices. This innovation, among others, is based on using a capacitator bank that dynamically adjusts the frequency. To validate the novel system, a series of experiments were conducted using commercial magnetic nanoparticles (MNPs) demonstrating the device’s effectiveness and allowing us to identify new challenges associated with the design of more powerful devices. A computational model was also used to validate the device and to allow us to determine the best system configuration. The results obtained are consistent with those from other studies using the same MNPs but with magnetic hyperthermia commercial equipment, confirming the good performance of the developed device (e.g., consistent SAR values between 1.37 and 10.80 W/gMNP were obtained, and experiments reaching temperatures above 43 °C were also obtained). This equipment offers additional advantages, including being economical, user-friendly, and portable. Full article
(This article belongs to the Special Issue New Insights into the Therapeutic Efficacy of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Three-dimensional view of solenoid S5. (<b>a</b>) The copper coil contains interior water for cooling inside, while the surrounding domain contains air (<b>b</b>) computational mesh.</p>
Full article ">Figure 2
<p>Initial slope method (ISM) for calculating the specific absorption rate (SAR).</p>
Full article ">Figure 3
<p>Parallel LC resonant circuit.</p>
Full article ">Figure 4
<p>Image of the switch set that controls the capacitor bank.</p>
Full article ">Figure 5
<p>(<b>a</b>) Initial prototype of our system, (<b>b</b>) the system being currently applied, (<b>c</b>) a new and more powerful prototype system (resonant) that is being tested.</p>
Full article ">Figure 6
<p>Overview of the system (<b>a</b>) and photographs of the apparatus and of the system (<b>b</b>): A—power source; B—system; C—solenoid; D—cooling system; E—equipment for measuring temperature; F—computer for acquiring and registering the temperature; G and H—oscilloscope coupled with a probe for measuring the magnetic field and verifying the waveform passing through the coil [<a href="#B39-nanomaterials-14-01848" class="html-bibr">39</a>].</p>
Full article ">Figure 6 Cont.
<p>Overview of the system (<b>a</b>) and photographs of the apparatus and of the system (<b>b</b>): A—power source; B—system; C—solenoid; D—cooling system; E—equipment for measuring temperature; F—computer for acquiring and registering the temperature; G and H—oscilloscope coupled with a probe for measuring the magnetic field and verifying the waveform passing through the coil [<a href="#B39-nanomaterials-14-01848" class="html-bibr">39</a>].</p>
Full article ">Figure 7
<p>Insulating system.</p>
Full article ">Figure 8
<p>Images of solenoids described in <a href="#nanomaterials-14-01848-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 9
<p>Profile of the magnetic field density (mT) of solenoid coil S5 for a driving voltage of 60 V and an operating frequency of 72 kHz. (<b>a</b>) Three-dimensional view, (<b>b</b>) two-dimensional view for the plane x = 0.</p>
Full article ">Figure 10
<p>Magnetic field density (mT) along the center of the simulated coils at various operating frequencies.</p>
Full article ">Figure 11
<p>The temperature variation of dstillated water samples was consistent across three experiments conducted on different days at a frequency of 69 kHz with solenoid S5.</p>
Full article ">Figure 12
<p>Heating curves for the Fluidmag<sub>ARA</sub> and Fluidmag<sub>UCA</sub> samples at a frequency 98 kHz using solenoid S5.</p>
Full article ">Figure 13
<p>Heating curves of Fluidmag<sub>D100nm</sub> (25 mg/mL) at 69 kHz using solenoid S5.</p>
Full article ">Figure 14
<p>Heating curves of Fluidmag<sub>D50nm</sub> (25 mg/mL) at 69 kHz using solenoid S5.</p>
Full article ">Figure 15
<p>Heating curves for the FluidMag<sub>D100nm</sub> sample at various frequencies using solenoid S5.</p>
Full article ">Figure 16
<p>Heating curves of various samples at a frequency of 138 kHz.</p>
Full article ">Figure 17
<p>Heating curves of FluidMag<sub>Dx100nn</sub> at different frequencies with a concentration of 12.5 mg<sub>MNP</sub>/mL using solenoid S6.</p>
Full article ">Figure 18
<p>Heating curves of FluidMag<sub>UCA</sub> at 72 kHz with a concentration of 12.5 mg<sub>MNP</sub>/mL using solenoid S8.</p>
Full article ">Figure 19
<p>Heating curves of FluidMag<sub>D100nm</sub> at different frequencies with a concentration of 12.5 mg<sub>MNP</sub>/mL using solenoid S8.</p>
Full article ">Figure 20
<p>Heating curves of FluidMag<sub>Dx50nm</sub> at different frequencies with a concentration of 25 mg<sub>MNP</sub>/mL using solenoid S8.</p>
Full article ">Figure 21
<p>Heating curves of FluidmagDX<sub>50nm</sub> samples with different concentrations (5 and 25 mg/m<sub>lMNP</sub>) at 132 kHz using solenoid S8.</p>
Full article ">Figure 22
<p>Heating and cooling curves of FluidMagDx50nm with the final configuration (S8).</p>
Full article ">
28 pages, 1948 KiB  
Review
Nanomaterial-Enhanced Hybrid Disinfection: A Solution to Combat Multidrug-Resistant Bacteria and Antibiotic Resistance Genes in Wastewater
by Tapas Kumar Mandal
Nanomaterials 2024, 14(22), 1847; https://doi.org/10.3390/nano14221847 - 19 Nov 2024
Abstract
This review explores the potential of nanomaterial-enhanced hybrid disinfection methods as effective strategies for addressing the growing challenge of multidrug-resistant (MDR) bacteria and antibiotic resistance genes (ARGs) in wastewater treatment. By integrating hybrid nanocomposites and nanomaterials, natural biocides such as terpenes, and ultrasonication, [...] Read more.
This review explores the potential of nanomaterial-enhanced hybrid disinfection methods as effective strategies for addressing the growing challenge of multidrug-resistant (MDR) bacteria and antibiotic resistance genes (ARGs) in wastewater treatment. By integrating hybrid nanocomposites and nanomaterials, natural biocides such as terpenes, and ultrasonication, this approach significantly enhances disinfection efficiency compared to conventional methods. The review highlights the mechanisms through which hybrid nanocomposites and nanomaterials generate reactive oxygen species (ROS) under blue LED irradiation, effectively disrupting MDR bacteria while improving the efficacy of natural biocides through synergistic interactions. Additionally, the review examines critical operational parameters—such as light intensity, catalyst dosage, and ultrasonication power—that optimize treatment outcomes and ensure the reusability of hybrid nanocomposites and other nanomaterials without significant loss of photocatalytic activity. Furthermore, this hybrid method shows promise in degrading ARGs, thereby addressing both microbial and genetic pollution. Overall, this review underscores the need for innovative wastewater treatment solutions that are efficient, sustainable, and scalable, contributing to the global fight against antimicrobial resistance. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic illustration depicting the limitations of conventional bacterial disinfection methods and the necessity for innovative disinfection strategies. Reprinted with permission from Ref. [<a href="#B23-nanomaterials-14-01847" class="html-bibr">23</a>] CC by 4.0.</p>
Full article ">Figure 2
<p>The diagram depicts the primary mechanisms involved in electro-oxidation, Electro-Fenton processes, ozone oxidation, and photocatalysis, all integrated with electrocoagulation for improved elimination of contaminants of emerging concern (CECs). Reprinted with permission from Ref. [<a href="#B44-nanomaterials-14-01847" class="html-bibr">44</a>] CC by 4.0.</p>
Full article ">Figure 3
<p>(<b>a</b>) Impact of radical scavengers on the photodegradation of methylene blue (MB) and (<b>b</b>) a schematic representation illustrating the photocatalytic mechanism of the NT/TiO<sub>2</sub> photocatalyst. Reprinted with permission from Ref. [<a href="#B102-nanomaterials-14-01847" class="html-bibr">102</a>], CC by 4.0.</p>
Full article ">Figure 4
<p>Mechanism of tetracycline degradation via photo-Fenton-like reaction. Reprinted with permission from Ref. [<a href="#B103-nanomaterials-14-01847" class="html-bibr">103</a>].</p>
Full article ">
16 pages, 4681 KiB  
Article
M-Doped (M = Zn, Mn, Ni) Co-MOF-Derived Transition Metal Oxide Nanosheets on Carbon Fibers for Energy Storage Applications
by Andrés González-Banciella, David Martinez-Diaz, Adrián de Hita, María Sánchez and Alejandro Ureña
Nanomaterials 2024, 14(22), 1846; https://doi.org/10.3390/nano14221846 - 19 Nov 2024
Abstract
Carbon fiber, with its strong mechanical properties and electrical conductivity, is ideal as a fiber electrode in wearable or structural energy storage devices. However, its energy storage capacity is limited, and coatings like transition metal oxides (TMOs) enhance its electrochemical performance. Metal–organic frameworks [...] Read more.
Carbon fiber, with its strong mechanical properties and electrical conductivity, is ideal as a fiber electrode in wearable or structural energy storage devices. However, its energy storage capacity is limited, and coatings like transition metal oxides (TMOs) enhance its electrochemical performance. Metal–organic frameworks (MOFs) are commonly used to grow TMOs on carbon fibers, increasing the surface area for better energy storage. Despite this, TMOs have limited electrical conductivity, so ion exchange is often used to dope them with additional cations, improving both conductivity and energy storage capacity. This study compares different ion-exchange cations in ZIF-L-derived TMO coatings on carbon fiber. Testing both supercapacitor and Li-ion battery applications, Ni-doped samples showed superior results, attributed to their higher exchange ratio with cobalt. As a supercapacitor electrode, the Ni-doped material achieved 13.3 F/g at 50 mA/g—66% higher than undoped samples. For Li-ion battery anodes, it reached a specific capacity of 410.5 mAh/g at 25 mA/g, outperforming undoped samples by 21.4%. Full article
(This article belongs to the Special Issue Metal Organic Framework (MOF)-Based Micro/Nanoscale Materials)
Show Figures

Figure 1

Figure 1
<p>SEM images of ZIF-L coating on carbon fiber fabric at (<b>a</b>) lower magnification and (<b>b</b>) higher magnification, (<b>c</b>) undoped TMO sample, (<b>d</b>) Zn-doped TMO sample, (<b>e</b>) Mn-doped TMO sample, and (<b>f</b>) Ni-doped TMO sample.</p>
Full article ">Figure 2
<p>(<b>a</b>) XPS surveys. (<b>b</b>) XPS Co 2p core level of the undoped sample. (<b>c</b>) XPS Zn 2p core level of Zn-doped sample. (<b>d</b>) XPS Mn 2p core level of Mn-doped sample. (<b>e</b>) XPS Ni 2p core level of Ni-doped sample. (<b>f</b>) The atomic percentage of each cation in each sample. (<b>g</b>) XRD spectra of TMO powder samples.</p>
Full article ">Figure 3
<p>TEM images of undoped TMO samples at (<b>a</b>) lower and (<b>b</b>) higher magnification.</p>
Full article ">Figure 4
<p>(<b>a</b>) CV curves at a scan rate of 20 mV/s. (<b>b</b>) Specific capacitance values are calculated from CV at 20 mV/s. (<b>c</b>) GCD test discharges at 50 mA/g. (<b>d</b>) Specific capacitance calculated from GCD tests at different current densities.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curve of Ni-doped sample at different scan rates and (<b>b</b>) calculated <span class="html-italic">b</span> value in the inset. (<b>c</b>) Diffusion and capacitive-controlled processes contribute to different scan rates for Ni-doped samples. (<b>d</b>) Specific capacitance and coulombic efficiency of Ni-doped sample after 5000 GCD cycles at 150 mA/g.</p>
Full article ">Figure 6
<p>(<b>a</b>) Normalized third-cycle CV of the samples. (<b>b</b>) GCD curves at 25 mA/g of the samples. (<b>c</b>) GCD calculated capacities at different current densities. (<b>d</b>) Nyquist plots of the samples.</p>
Full article ">Figure 7
<p>(<b>a</b>) Ni-doped sample CV at different scan rates and calculated b value in the inset. (<b>b</b>) Percentage of each contribution at different scan rates in the Ni-doped sample. (<b>c</b>) Specific capacity of Ni-doped sample after 100 GCD cycles at 100 mA/g. (<b>d</b>) Nyquist plot of Ni-doped sample after several number of cycles.</p>
Full article ">
5 pages, 198 KiB  
Editorial
Micro- and Nanostructured Biomaterials for Biomedical Applications and Regenerative Medicine
by Michele Bianchi and Gianluca Carnevale
Nanomaterials 2024, 14(22), 1845; https://doi.org/10.3390/nano14221845 - 18 Nov 2024
Viewed by 328
Abstract
Over the past two decades, research on innovative micro- and nano-biomaterials has seen a significant surge in the bioengineering, biomedicine, and regenerative medicine fields [...] Full article
13 pages, 3286 KiB  
Article
Improving the NO2 Gas Sensing Performances at Room Temperature Based on TiO2 NTs/rGO Heterojunction Nanocomposites
by Yan Ling, Yunjiang Yu, Canxin Tian and Changwei Zou
Nanomaterials 2024, 14(22), 1844; https://doi.org/10.3390/nano14221844 - 18 Nov 2024
Viewed by 323
Abstract
The development of energy-efficient, sensitive, and reliable gas sensors for monitoring NO2 concentrations has garnered considerable attention in recent years. In this manuscript, TiO2 nanotube arrays/reduced graphene oxide nanocomposites with varying rGO contents (TiO2 NTs/rGO) were synthesized via a two-step [...] Read more.
The development of energy-efficient, sensitive, and reliable gas sensors for monitoring NO2 concentrations has garnered considerable attention in recent years. In this manuscript, TiO2 nanotube arrays/reduced graphene oxide nanocomposites with varying rGO contents (TiO2 NTs/rGO) were synthesized via a two-step method for room temperature NO2 gas detection. From SEM and TEM images, it is evident that the rGO sheets not only partially surround the TiO2 nanotubes but also establish interconnection bridges between adjacent nanotubes, which is anticipated to enhance electron–hole separation by facilitating electron transfer. The optimized TiO2 NTs/rGO sensor demonstrated a sensitive response of 19.1 to 1 ppm of NO2, a 5.26-fold improvement over the undoped TiO2 sensor. Additionally, rGO doping significantly enhanced the sensor’s response/recovery times, reducing them from 24 s/42 s to 18 s/33 s with just 1 wt.% rGO. These enhancements are attributed to the increased specific surface area, higher concentration of chemisorbed oxygen species, and the formation of p-n heterojunctions between TiO2 and rGO within the nanocomposites. This study provides valuable insights for the development of TiO2/graphene-based gas sensors for detecting oxidizing gases at room temperature. Full article
(This article belongs to the Special Issue Design and Applications of Heterogeneous Nanostructured Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic of synthesis process (<b>a</b>) and photograph of NO<sub>2</sub> detection testing system (<b>b</b>).</p>
Full article ">Figure 2
<p>XRD patterns of TiO<sub>2</sub> NTs/rGO nanocomposites grown with different rGO contents.</p>
Full article ">Figure 3
<p>SEM images of TiO<sub>2</sub> NTs/rGO nanocomposites with rGO contents of 0 wt.% (<b>a</b>), 0.5 wt.% (<b>b</b>), 1 wt.% (<b>c</b>), and 3 wt.% (<b>d</b>), respectively. Inset of <a href="#nanomaterials-14-01844-f003" class="html-fig">Figure 3</a>d shows the EDX spectrum of TiO<sub>2</sub> NTs/rGO nanocomposites with rGO contents of 3 wt.%. The red circles in <a href="#nanomaterials-14-01844-f003" class="html-fig">Figure 3</a>d illustrate the structure of rGO surrounding the TiO<sub>2</sub> nanotubes.</p>
Full article ">Figure 4
<p>TEM (<b>a</b>,<b>b</b>) and high-resolution TEM (<b>c</b>,<b>d</b>) images of TiO<sub>2</sub> nanotubes (<b>a</b>,<b>c</b>) and TiO<sub>2</sub> NTs/rGO nanocomposites (<b>b</b>,<b>d</b>) with rGO contents of 1 wt.%.</p>
Full article ">Figure 5
<p>Raman spectra of TiO2 NTs/rGO nanocomposites with different rGO contents.</p>
Full article ">Figure 6
<p>XPS spectra of TiO<sub>2</sub> NTs/rGO nanocomposite with rGO contents of 1 wt.%. (<b>a</b>) Full scan. (<b>b</b>) Ti 2p. (<b>c</b>) C 1s. (<b>d</b>) O 1s.</p>
Full article ">Figure 7
<p>The response value of the sensor based on TiO<sub>2</sub> NTs/rGO composites vs. NO<sub>2</sub> concentration at room temperature.</p>
Full article ">Figure 8
<p>The dynamic response transients of the sensor based on TiO<sub>2</sub> NTs/rGO nanocomposites to 20 ppm NO<sub>2</sub> at room temperature.</p>
Full article ">Figure 9
<p>Response of the sensor based on TiO<sub>2</sub> NTs/rGO nanocomposites to 50 ppm of C<sub>2</sub>H<sub>5</sub>OH, CH<sub>3</sub>OH, H<sub>2</sub>, NH<sub>3</sub>, H<sub>2</sub>S, and NO<sub>2</sub> at room temperature.</p>
Full article ">Figure 10
<p>Long-term stability of the sensor based on TiO<sub>2</sub> NTs and TiO<sub>2</sub> NTs/rGO nanocomposites to 1 ppm of NO<sub>2</sub> at room temperature.</p>
Full article ">Figure 11
<p>I-V curves of the sensor based on TiO<sub>2</sub> NTs/rGO nanocomposites with different rGO contents.</p>
Full article ">Figure 12
<p>Energy band diagrams for TiO<sub>2</sub> NTs, rGO and TiO<sub>2</sub> NTs/rGO heterostructure, where E<sub>VB</sub>, E<sub>F</sub>, E<sub>CB</sub>, E<sub>vac</sub> represent valence band, Fermi level, conduction band, and vacuum level, respectively. (<b>a</b>) The energy band diagrams for TiO<sub>2</sub> nanotubes, rGO, and their corresponding TiO<sub>2</sub> NTs/rGO nanocomposites. (<b>b</b>) Schematic illustration of electron transfer and sensing mechanism of TiO<sub>2</sub> NTs/rGO nanocomposites.</p>
Full article ">
12 pages, 5364 KiB  
Article
Controlled Formation of Silicon-Vacancy Centers in High-Pressure Nanodiamonds Produced from an “Adamantane + Detonation Nanodiamond” Mixture
by Dmitrii G. Pasternak, Rustem H. Bagramov, Alexey M. Romshin, Igor P. Zibrov, Vladimir P. Filonenko and Igor I. Vlasov
Nanomaterials 2024, 14(22), 1843; https://doi.org/10.3390/nano14221843 - 18 Nov 2024
Viewed by 338
Abstract
Despite progress in the high-pressure synthesis of nanodiamonds from hydrocarbons, the problem of controlled formation of fluorescent impurity centers in them still remains unresolved. In our work, we explore the potential of a new precursor composition, a mixture of adamantane with detonation nanodiamond, [...] Read more.
Despite progress in the high-pressure synthesis of nanodiamonds from hydrocarbons, the problem of controlled formation of fluorescent impurity centers in them still remains unresolved. In our work, we explore the potential of a new precursor composition, a mixture of adamantane with detonation nanodiamond, both in the synthesis of nanodiamonds and in the controlled formation of negatively charged silicon-vacancy centers in such nanodiamonds. Using different adamantane/detonation nanodiamond weight ratios, a series of samples was synthesized at a pressure of 7.5 GPa in the temperature range of 1200–1500 °C. It was found that temperature around 1350 °C, is optimal for the high-yield synthesis of nanodiamonds <50 nm in size. For the first time, controlled formation of negatively charged silicon-vacancy centers in such small nanodiamonds was demonstrated by varying the atomic ratios of silicon/carbon in the precursor in the range of 0.01–1%. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD patterns of four different HPHT samples synthesized at 1250 °C from adamantane/DND mixture in the following weight proportions: 0/1, 1/1, 10/1 and 100/1. For convenience of comparative analysis, the diffraction patterns are normalized to the intensity of the (220) diamond diffraction peak, and the background associated with the scattering of X-rays on the substrate is subtracted from them. D—diamond, G—graphite. (<b>b</b>) Zoomed part of XRD pattern around (220) diamond diffraction peak.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM images of samples synthesized at 1250 °C from adamantane/DND mixture in weight proportion 10/1 (<b>a</b>) and 100/1 (<b>b</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) Wide-angle XRD patterns of “1/1” (black) and “25/1” (red) samples synthesized at 1350 and 1450 °C. (<b>b</b>) Approximation of (220) diamond diffraction peaks by Lorentz profiles for samples “1/1” and “25/1” synthesized at 1350 °C. The peak shape is most accurately fitted by the superposition of three profiles. The arrows indicate the average sizes of diamond nanoparticles estimated for each profile.</p>
Full article ">Figure 4
<p>SEM, TEM images and Raman spectrum of the sample “25/1” synthesized at 1350 °C. (<b>a</b>) SEM image of diamond nanoparticles dispersed on a substrate from alcohol suspension. (<b>b</b>) TEM image of the ND agglomerate prepared by drying a drop of the suspension on a grid. The characteristic sizes of diamond crystallites are less than 10 nm in the marked area A, and 20–50 nm in the area B. (<b>c</b>) Typical Raman spectrum of the sample. The diamond peak (orange dotted line) is shifted to the position of 1332.1 cm<sup>−</sup><sup>1</sup> and broadened up to 5.7 cm<sup>−1</sup> relative to the diamond peak (black line) recorded for the natural bulk IIa diamond.</p>
Full article ">Figure 5
<p>PL spectra of the samples synthesized at 1350 °C from adamantane/DND/tetrakis mixture at adamantane/DND weight proportions 25/1 (red) and 1/1 (black), and Si/C atomic ratio 1%. The spectra are normalized to the intensity of the diamond Raman (DR) line. Zero-phonon line of SiV<sup>−</sup> center and DR line are observed at 738.5 nm and 504.8 nm, respectively. “Palisade” of narrow lines observed throughout the range 500–800 nm is typical for N-doped H-terminated NDs [<a href="#B43-nanomaterials-14-01843" class="html-bibr">43</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) PL spectra of the sample “25/1” synthesized at 1350 °C from adamantane/DND/tetrakis mixture with atomic ratio Si/C varying from 0% to 1%. The spectra are normalized to the intensity of the diamond Raman (DR) line. (<b>b</b>) The dependence of the integrated fluorescence intensity SiV<sup>−</sup> normalized to the DR line (I<sup>int</sup><sub>SiV/DR</sub>) on the atomic ratio Si/C in the precursor. The statistical error in determining I<sup>int</sup><sub>SiV/DR</sub>/DR was calculated based on averaging ten PL spectra for each of the five samples.</p>
Full article ">Figure 7
<p>Raman spectrum of the sample synthesized at 1350 °C from chloradamantane/tetrakis mixture with atomic ratio Si/C = 1/1000. Only the lines of DR (504.8 nm) and the stretching vibrations of surface CHx groups (near 550 nm) are observed in the spectrum.</p>
Full article ">Figure A1
<p>XRD pattern of three samples: initial (black), treated at 7.5 GPa pressure and temperatures of 1250 °C (orange) and 1450 °C (red). Only diffraction peaks of diamond are observed in all three samples. The estimation of the crystallite sizes according to the Scherrer formula indicates that the average size of diamond crystallites increases from 4–5 nm (initial) to 7 nm (1450 °C). Note that amorphous carbon, which dominates in the Raman spectra (see below), is not detected by the XRD.</p>
Full article ">Figure A2
<p>(<b>a</b>) The typical Raman spectrum of the untreated DND consists of an asymmetric diamond peak (DR) at 1328.5 cm<sup>−1</sup> and a band at 1630 cm<sup>−1</sup> that overlap with the initial fragment of broadband (500–700 nm) fluorescence background [<a href="#B37-nanomaterials-14-01843" class="html-bibr">37</a>]. (<b>b</b>) Typical Raman spectrum of DND treated at 1450 °C is characterized by a broad band of amorphous carbon (1300–1600 cm<sup>−1</sup>), which dominates the spectrum [<a href="#B49-nanomaterials-14-01843" class="html-bibr">49</a>]; a diamond (DR) peak is barely distinguished against the fluorescence background. Similar Raman spectra were observed for DND treated at 1250 °C.</p>
Full article ">
13 pages, 2793 KiB  
Article
Nature of the Pits on the Lattice-Matched InAlAs Layer Surface Grown on the (001) InP Substrate
by Dmitrii V. Gulyaev, Demid S. Abramkin, Dmitriy V. Dmitriev, Alexander I. Toropov, Eugeniy A. Kolosovsky, Sergey A. Ponomarev, Nina N. Kurus, Ilya A. Milekhin and Konstantin S. Zhuravlev
Nanomaterials 2024, 14(22), 1842; https://doi.org/10.3390/nano14221842 - 18 Nov 2024
Viewed by 282
Abstract
The structural properties of lattice-matched InAlAs/InP layers grown by molecular beam epitaxy have been studied using atomic force microscopy, scanning electron microscopy and micro-photoluminescence spectroscopy. The formation of the surface pits with lateral sizes in the micron range and a depth of about [...] Read more.
The structural properties of lattice-matched InAlAs/InP layers grown by molecular beam epitaxy have been studied using atomic force microscopy, scanning electron microscopy and micro-photoluminescence spectroscopy. The formation of the surface pits with lateral sizes in the micron range and a depth of about 2 ÷ 10 nm has been detected. The InP substrate annealing temperature and value of InAlAs alloy composition deviation from the lattice-matched InxAl1−xAs/InP case (x = 0.52) control the density of pits ranging from 5 × 105 cm−2 ÷ 108 cm−2. The pit sizes are controlled by the InAlAs layer thickness and growth temperature. The correlation between the surface pits and threading dislocations has been detected. Moreover, the InAlAs surface is characterized by composition inhomogeneity with a magnitude of 0.7% with the cluster lateral sizes and density close to these parameters for surface pits. The experimental data allow us to suggest a model where the formation of surface pits and composition clusters is caused by the influence of a local strain field in the threading dislocation core vicinity on In adatoms incorporating kinetic. Full article
(This article belongs to the Section Nanophotonics Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>Scheme of structures with In<sub>x</sub>Al<sub>1−x</sub>As layer with thickness <span class="html-italic">D</span> grown at <span class="html-italic">T</span><sub>S</sub>. The preparation of the InP substrate was performed at <span class="html-italic">T</span><sub>A</sub>. The InAsP layer, formed during the substrate preparation, is marked as «InAsP».</p>
Full article ">Figure 2
<p>AFM images of 1 μm thick InAlAs layers lattice-matched with the substrate. <span class="html-italic">T</span><sub>A</sub> = 485 °C and <span class="html-italic">T</span><sub>S</sub> = 485 °C for (<b>A</b>) and <span class="html-italic">T</span><sub>A</sub> = 535 °C, <span class="html-italic">T</span><sub>S</sub> = 525 °C for (<b>B</b>). Total layer thickness is 1000 nm for both structures. Relief profiles are presented in the insets.</p>
Full article ">Figure 3
<p>(<b>a</b>) The dependence of the pit density on the substrate annealing temperature for the lattice-matched layer (<span class="html-italic">x</span> = 0.52, black dots) and on the layer composition deviation from the lattice-matched value (red dots). (<b>b</b>) The dependence of the lateral size of the pits on the InAlAs layer thickness for <span class="html-italic">x</span> = 0.52 (black dots) and <span class="html-italic">x</span> = 0.51 (red dots). (<b>c</b>) The dependence of the depth of the pits on the thickness of the InAlAs layer with <span class="html-italic">x</span> = 0.52 grown at the temperature of 505 °C. (<b>d</b>) The dependence of the depth of the pits on the growth temperature for the InAlAs layer with <span class="html-italic">x</span> = 0.52 grown at the total layer thickness of 1 μm.</p>
Full article ">Figure 4
<p>The SEM cross-sectional image of the heterostructure grown at <span class="html-italic">T</span><sub>A</sub> = 535 °C, <span class="html-italic">T</span><sub>S</sub> = 525 °C and <span class="html-italic">D</span> = 1000 nm, with the alloy composition deviation of 0.5%. The vertical arrows at the top panel point to the surface pits. The bottom panel shows the same area of the SEM image but with the dislocations indicated by thin white lines for better clarity.</p>
Full article ">Figure 5
<p>(<b>A</b>) A typical map of the position of the PL peak of the InAlAs layer with pits. (<b>B</b>) Micro-PL spectra from the InAlAs layer measured (1) inside and (outside) the cluster. The annealing temperature of the substrate/growth is 535/505 °C.</p>
Full article ">Figure 6
<p>Experimental and calculated temperature dependencies of the pit depth (black square) and In depletion (red circle) for lattice-matched 1 μm thick InAlAs layers.</p>
Full article ">
11 pages, 2655 KiB  
Article
Enhanced Optical and Electrical Properties of IGZO/Ag/IGZO for Solar Cell Application via Post-Rapid Thermal Annealing
by Chanmin Hwang, Taegi Kim, Yuseong Jang, Doowon Lee and Hee-Dong Kim
Nanomaterials 2024, 14(22), 1841; https://doi.org/10.3390/nano14221841 - 18 Nov 2024
Viewed by 295
Abstract
In this paper, we optimized IGZO/Ag/IGZO (IAI) multilayer films by post-rapid thermal annealing (RTA) to enhance the electrical conductivity and optical transmittance in visible wavelengths for solar cell applications. Our optimized device showed an average transmittance of 85% in the visible range, with [...] Read more.
In this paper, we optimized IGZO/Ag/IGZO (IAI) multilayer films by post-rapid thermal annealing (RTA) to enhance the electrical conductivity and optical transmittance in visible wavelengths for solar cell applications. Our optimized device showed an average transmittance of 85% in the visible range, with a lowest sheet resistance of 6.03 Ω/□ when annealed at 500 °C for 60 s. Based on these results, we assessed our device with photo-generated short circuit current density (JSC) using a solar cell simulator to confirm its applicability in the solar cell. IAI multilayer RTA at 500 °C for 60 s showed a highest JSC of 40.73 mA/cm2. These results show that our proposed IAI multilayer film, which showed a high optical transparency and electrical conductivity optimized with post RTA, seems to be excellent transparent electrode for solar cell applications. Full article
(This article belongs to the Special Issue Nanostructured Materials for Electric Applications)
Show Figures

Figure 1

Figure 1
<p>The RTA process for the IGZO/Ag/IGZO multilayer, and the right diagram describes the generated oxygen vacancies from the IGZO inside by RTA.</p>
Full article ">Figure 2
<p>AFM topology images of IGZO/Ag/IGZO after RTA, (<b>a</b>) RT, (<b>b</b>) 300, (<b>c</b>) 400, and (<b>d</b>) 500 °C for 60 s.</p>
Full article ">Figure 3
<p>The sheet resistance of (<b>a</b>) IGZO and (<b>b</b>) IGZO/Ag/IGZO multilayers after varying temperature conditions.</p>
Full article ">Figure 4
<p>(<b>a</b>) Impedance of IGZO/Ag/IGZO after RTA RT to 500 °C for 60 s and (<b>b</b>) equivalent circuit of IGZO/Ag/IGZO.</p>
Full article ">Figure 5
<p>(<b>a</b>) Transmittance, (<b>b</b>) reflectivity, and (<b>c</b>) absorption coefficient of IGZO/Ag/IGZO after RTA RT to 500 °C for 60 s, and each inset depicts the transmittance and reflectivity at 500 nm wavelength with varying annealing temperature.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic structure of IGZO/Ag/IGZO-based solar cell and (<b>b</b>) the photo-generated short circuit current of IGZO/Ag/IGZO-based solar cell as a function of annealing temperature.</p>
Full article ">
18 pages, 2873 KiB  
Article
Improving Resistive Heating, Electrical and Thermal Properties of Graphene-Based Poly(Vinylidene Fluoride) Nanocomposites by Controlled 3D Printing
by Rumiana Kotsilkova, Vladimir Georgiev, Mariya Aleksandrova, Todor Batakliev, Evgeni Ivanov, Giovanni Spinelli, Rade Tomov and Tsvetozar Tsanev
Nanomaterials 2024, 14(22), 1840; https://doi.org/10.3390/nano14221840 - 17 Nov 2024
Viewed by 517
Abstract
This study developed a novel 3D-printable poly(vinylidene fluoride) (PVDF)-based nanocomposite incorporating 6 wt% graphene nanoplatelets (GNPs) with programmable characteristics for resistive heating applications. The results highlighted the significant effect of a controlled printing direction (longitudinal, diagonal, and transverse) on the electrical, thermal, Joule [...] Read more.
This study developed a novel 3D-printable poly(vinylidene fluoride) (PVDF)-based nanocomposite incorporating 6 wt% graphene nanoplatelets (GNPs) with programmable characteristics for resistive heating applications. The results highlighted the significant effect of a controlled printing direction (longitudinal, diagonal, and transverse) on the electrical, thermal, Joule heating, and thermo-resistive properties of the printed structures. The 6 wt% GNP/PVDF nanocomposite exhibited a high electrical conductivity of 112 S·m−1 when printed in a longitudinal direction, which decreased significantly in other directions. The Joule heating tests confirmed the material’s efficiency in resistive heating, with the maximum temperature reaching up to 65 °C under an applied low voltage of 2 V at a raster angle of printing of 0°, while the heating Tmax decreased stepwise with 10 °C at the 45° and the 90° printing directions. The repeatability of the Joule heating performance was verified through multiple heating and cooling cycles, demonstrating consistent maximum temperatures across several tests. The effect of sample thickness, controlled by the number of printed layers, was investigated, and the results underscore the advantages of programmable 3D printing orientation in thin layers for enhanced thermal stability, tailored electrical conductivity, and efficient Joule heating capabilities of 6 wt% GNP/PVDF composites, positioning them as promising candidates for next-generation 3D-printed electronic devices and self-heating applications. Full article
(This article belongs to the Special Issue Hybrid Nano Polymer Composites (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>TEM images: (<b>a</b>) GNP filler surfaces with SEAD pattern (inset); (<b>b</b>) high-resolution TEM image of the GNP thickness showing the multi-layered structure of oriented graphene monolayers; and (<b>c</b>) exfoliated GNP nanostructures dispersed in the PVDF matrix. Arrows show the thickness of the exfoliated GNPs.</p>
Full article ">Figure 2
<p>Thermal properties of PVDF and 6 wt% GNP/PVDF: DSC thermograms of heat flow vs. temperature at a scan rate of 10 °C/min, showing the first heating run (<b>a</b>), cooling cycle (<b>b</b>), and second heating run (<b>c</b>). The dash lines point the thermal transitions of the neat PVDF. In (<b>d</b>), the TGA/DTG thermograms of mass loss vs. temperature for the polymer and the nanocomposite are plotted, while the GNP thermogram is presented in the inset figure.</p>
Full article ">Figure 3
<p>SEM micrographs of the cut surface of samples with different deposition directions: (<b>a</b>) longitudinal (3DP 0°); (<b>b</b>) diagonal (3DP 45°); (<b>c</b>) transverse (3DP 90°); and (<b>d</b>) voltage vs. current dependence, varying the printing directions. The magnification bar is 1 mm. The arrows show the current flow direction.</p>
Full article ">Figure 4
<p>Comparison of (<b>a</b>) temperature vs. time and (<b>b</b>) temperature increase and heat vs. electrical conductivity of 6 wt% GNP/PVDF, varying the 3D printing directions—3DP 0°, 3DP 45°, and 3DP 90°—for 2 mm thick samples at an applied voltage of 2 V.</p>
Full article ">Figure 5
<p>Temperature vs. time for four-cycle heating–cooling test of 6 wt% GNP/PVDF samples at an applied voltage of 2 V with various printing directions: (<b>a</b>) longitudinal 3DP 0°, (<b>b</b>) diagonal 3DP 45°, and (<b>c</b>) transverse 3DP 90°. (<b>d</b>) Repeatability of the maximal temperature and current in the four heating–cooling cycles for the three printing directions.</p>
Full article ">Figure 6
<p>Temperature and current vs. time for 6 wt% GNP/PVDF, for the diagonally printed samples (3DP 45°) with (<b>a</b>) 4 printed layers (0.8 mm thick) and (<b>b</b>) 10 printed layers (2 mm thick), varying the applied voltage.</p>
Full article ">Figure 7
<p>Comparison of (<b>a</b>) maximum heating temperature and current vs. applied voltage and (<b>b</b>) generated heat and heating efficiency vs. power for the 3DP45° samples of the 6 wt% GNP/PVDF nanocomposite, with a controlled number of printed layers (4 layers, 0.8 mm thick; and 10 layers, 2 mm thick).</p>
Full article ">Figure 8
<p>Resistance vs. temperature of the 6 wt% GNP/PVDF composites with various printing directions of 3DP 0°, 3DP 45°, and 3DP 90°.</p>
Full article ">Figure 9
<p>Thermal diffusivity and conductivity of the 3DP samples of 6 wt% GNP/PVDF vs. temperature, with various printing directions.</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 330
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 ">
12 pages, 5559 KiB  
Article
Potassium-Based Solid Sorbents for CO2 Adsorption: Key Role of Interconnected Pores
by Yuan Zhao, Jiangbo Huo, Xuefei Wang and Shunwei Ma
Nanomaterials 2024, 14(22), 1838; https://doi.org/10.3390/nano14221838 - 17 Nov 2024
Viewed by 287
Abstract
Industrial CO2 emissions contribute to pollution and greenhouse effects, highlighting the importance of carbon capture. Potassium carbonate (K2CO3) is an effective CO2 absorbent, yet its liquid-phase absorption faces issues like diffusion resistance and corrosion risks. In this [...] Read more.
Industrial CO2 emissions contribute to pollution and greenhouse effects, highlighting the importance of carbon capture. Potassium carbonate (K2CO3) is an effective CO2 absorbent, yet its liquid-phase absorption faces issues like diffusion resistance and corrosion risks. In this work, the solid adsorbents were developed with K2CO3 immobilized on the selected porous supports. Al2O3 had an optimum CO2 adsorption capacity of 0.82 mmol g−1. After further optimization of its pore structure, the self-prepared support Al2O3-2, which has an average pore diameter of 11.89 nm and a pore volume of 0.59 cm3 g−1, achieved a maximum CO2 adsorption capacity of 1.12 mmol g−1 following K2CO3 impregnation. Additionally, the relationship between support structure and CO2 adsorption efficiency was also analyzed. The connectivity of the pores and the large pore diameter of the support may play a key role in enhancing CO2 adsorption performance. During 10 cycles of testing, the K2CO3-based adsorbents demonstrated consistent high CO2 adsorption capacity with negligible degradation. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the experimental setup for CO<sub>2</sub> adsorption assessment.</p>
Full article ">Figure 2
<p>CO<sub>2</sub> adsorption capacities of the 20% K<sub>2</sub>CO<sub>3</sub>/supports under the conditions: 10 vol% CO<sub>2</sub> in air, 298 K adsorption temperature.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) CO<sub>2</sub> adsorption capacities of 20% K<sub>2</sub>CO<sub>3</sub>/Al<sub>2</sub>O<sub>3</sub>-2 adsorbents by varying preparation conditions.</p>
Full article ">Figure 4
<p>XRD patterns of Al<sub>2</sub>O<sub>3</sub>-1 and Al<sub>2</sub>O<sub>3</sub>-2.</p>
Full article ">Figure 5
<p>SEM images of Al<sub>2</sub>O<sub>3</sub>-1 (<b>a</b>,<b>b</b>) and Al<sub>2</sub>O<sub>3</sub>-2 (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 6
<p>N<sub>2</sub> adsorption (hollow) and desorption (solid) isotherms of Al<sub>2</sub>O<sub>3</sub>-1 and Al<sub>2</sub>O<sub>3</sub>-2.</p>
Full article ">Figure 7
<p>Pore size distributions of Al<sub>2</sub>O<sub>3</sub>-1 and Al<sub>2</sub>O<sub>3</sub>-2.</p>
Full article ">Figure 8
<p>A comparative schematic of CO<sub>2</sub> diffusion in interconnected and non-interconnected pore channels. The arrow represents the direction of molecular diffusion.</p>
Full article ">Figure 9
<p>TGA curves of Al<sub>2</sub>O<sub>3</sub>-1 and Al<sub>2</sub>O<sub>3</sub>-2.</p>
Full article ">Figure 10
<p>Recycling performance of 20% K<sub>2</sub>CO<sub>3</sub>/Al<sub>2</sub>O<sub>3</sub>-2 adsorbent after 10 cycles.</p>
Full article ">
20 pages, 6771 KiB  
Article
Enhancement of the Physical and Mechanical Properties of Cellulose Nanofibril-Reinforced Lignocellulosic Foams for Packaging and Building Applications
by Mara Paulette Alonso, Rakibul Hossain, Maryam El Hajam and Mehdi Tajvidi
Nanomaterials 2024, 14(22), 1837; https://doi.org/10.3390/nano14221837 - 17 Nov 2024
Viewed by 462
Abstract
Biobased foams have the potential to serve as eco-friendly alternatives to petroleum-based foams, provided they achieve comparable thermomechanical and physical properties. We propose a facile approach to fabricate eco-friendly cellulose nanofibril (CNF)-reinforced thermomechanical pulp (TMP) fiber-based foams via an oven-drying process with thermal [...] Read more.
Biobased foams have the potential to serve as eco-friendly alternatives to petroleum-based foams, provided they achieve comparable thermomechanical and physical properties. We propose a facile approach to fabricate eco-friendly cellulose nanofibril (CNF)-reinforced thermomechanical pulp (TMP) fiber-based foams via an oven-drying process with thermal conductivity as low as 0.036 W/(m·K) at a 34.4 kg/m3 density. Acrodur®, iron chloride (FeCl3), and cationic polyacrylamide (CPAM) were used to improve the foam properties. Acrodur® did not have any significant effect on the foamability and density of the foams. Mechanical, thermal, cushioning, and water absorption properties of the foams were dependent on the density and interactions of the additives with the fibers. Due to their high density, foams with CPAM and FeCl3 at a 1% additive dosage had significantly higher compressive properties at the expense of slightly higher thermal conductivity. There was slight increase in compressive properties with the addition of Acrodur®. All additives improved the water stability of the foams, rendering them stable even after 24 h of water absorption. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic representation of the fabrication process of surfactant-assisted cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams.</p>
Full article ">Figure 2
<p>Stability of the aqueous foams (<b>a</b>) without any additive (neat foams), and with (<b>b</b>) 1% Acrodur<sup>®</sup>, (<b>c</b>) 1% CPAM, and (<b>d</b>) 1% FeCl<sub>3</sub> as additives. All additives were added at 1% of total fibers (based on dry weight). When fibers were used, the composition was 95% TMP fibers and 5% CNFs.</p>
Full article ">Figure 3
<p>Foamability values before and after adding thermomechanical pulp (TMP) fibers during gravity filtration for different formulations for the preparation of the cellulose nanofibril-reinforced TMP fiber-based foams.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>g</b>) Digital photographs of the cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams of different formulations and their respective thicknesses.</p>
Full article ">Figure 5
<p>(<b>a</b>) Density and (<b>b</b>) porosity of the cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams of different formulations. Values with common letters are not significantly different from each other at a significant level of 0.05.</p>
Full article ">Figure 6
<p>Scanning electron microscopy (SEM) images of the surfaces of the cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams with (<b>a</b>) no additives, (<b>b</b>) 1% Acrodur<sup>®</sup>, (<b>c</b>) 1% FeCl<sub>3</sub>, and (<b>d</b>) 1% CPAM as additives.</p>
Full article ">Figure 7
<p>Energy-dispersive X-ray spectroscopy (EDS) images of the surface of the cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams with (<b>a</b>) 0.5% FeCl<sub>3</sub> (<b>b</b>) 1% FeCl<sub>3</sub>, (<b>c</b>) 0.5% CPAM, and (<b>d</b>) 1% CPAM as additives.</p>
Full article ">Figure 8
<p>(<b>a</b>) Compressive strength at 10% and 25% strain, (<b>b</b>) compressive modulus, (<b>c</b>) thickness recovery after 35% compressive strain, and (<b>d</b>) foam resilience properties for cellulose nanofibrils-reinforced thermomechanical pulp fiber-based foams of different formulations. Values with common letters are not significantly different from each other at a significant level of 0.05.</p>
Full article ">Figure 9
<p>(<b>a</b>) Density, (<b>b</b>) tensile strain, normalized (<b>c</b>) tensile strength, and (<b>d</b>) tensile modulus of films (without SDS) with different additives. Values with common letters are not significantly different from each other at a significant level of 0.05.</p>
Full article ">Figure 10
<p>Scanning electron microscopy (SEM) images of the fractured surfaces of the cellulose nanofibrils-reinforced thermomechanical pulp fiber-based films with (<b>a</b>) no additive, (<b>b</b>) 1% Acrodur<sup>®</sup>, (<b>c</b>) 1% FeCl<sub>3</sub>, and (<b>d</b>) 1% CPAM as additives after tensile failure.</p>
Full article ">Figure 11
<p>(<b>a</b>) Water absorption (by volume) and (<b>b</b>) thickness swelling of the cellulose nanofibril-reinforced thermomechanical pulp fiber-based foams of different formulations for 2 h and 24 h test times. ‘X’ indicates disintegration of foams after water absorption. Values with common letters are not significantly different from each other at a significant level of 0.05.</p>
Full article ">Figure 12
<p>(<b>a</b>) Thermal conductivity and (<b>b</b>) the relationship between thermal conductivity and density of the cellulose nanofibrils-reinforced thermomechanical pulp fiber-based foams of different formulations. Values with common letters are not significantly different from each other at a significant level of 0.05.</p>
Full article ">
15 pages, 3359 KiB  
Article
Improvement in Curcumin’s Stability and Release by Formulation in Flexible Nano-Liposomes
by Hua-Wei Chen, Su-Der Chen, Hung-Ta Wu, Chun-Hung Cheng, Chyow-San Chiou and Wei-Ting Chen
Nanomaterials 2024, 14(22), 1836; https://doi.org/10.3390/nano14221836 - 17 Nov 2024
Viewed by 300
Abstract
Curcumin is utilized extensively as Chinese medicine in Asia due to its antioxidant, antimicrobial, and inflammatory activities. However, its use has the challenges of low oral bioavailability and high heat sensitivity. The aim of this research was to produce flexible nano-liposomes containing curcumin [...] Read more.
Curcumin is utilized extensively as Chinese medicine in Asia due to its antioxidant, antimicrobial, and inflammatory activities. However, its use has the challenges of low oral bioavailability and high heat sensitivity. The aim of this research was to produce flexible nano-liposomes containing curcumin using an innovative approach of ethanol injection and Tween 80 to enhance the stability and preservation of curcumin. The mean particle size, encapsulation efficiency, thermal degradation, storage stability, and curcumin release in flexible nano-liposomes were also investigated. We found that the mean particle size of curcumin-loaded flexible nano-liposome decreased from 278 nm to 27.6 nm. At the same time, the Tween 80 concentration increased from 0 to 0.15 wt%, which corresponded with the results of transmission electron microscopy (TEM) morphology analyses, and particle size decreased with an enhancement in Tween 80 concentration. Further, pure curcumin was quickly released within one hour at 37 °C, and first-order kinetics matched with its release curve. However, curcumin encapsulated in flexible nano-liposomes showed a slow release of 71.24% within 12 h, and a slower release pattern matched with the Higuchi model over 24 h, ultimately reaching 84.63% release. Hence, flexible nano-liposomes of curcumin made by a combination of ethanol injection and Tween 80 addition prevented the thermal degradation of curcumin, and enhanced its storage stability and preservation for future drug delivery applications. Full article
(This article belongs to the Special Issue Green Nanoparticles for Topical Administration of Drugs)
Show Figures

Figure 1

Figure 1
<p>Transmission electron microscopy (TEM) images of flexible nano-liposomes loaded with curcumin prepared with different Tween 80 concentrations (<b>a</b>) 0.00 wt% (<b>b</b>) 0.05 wt% (<b>c</b>) 0.15 wt% (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL).</p>
Full article ">Figure 2
<p>The effect of different Tween 80 concentrations on flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL).</p>
Full article ">Figure 3
<p>Zeta potential of various liposome contents with and without Tween 80 (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 4
<p>Degradation of curcumin at different temperatures with or without encapsulation in flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 5
<p>Effects of different temperatures on curcumin liposomes and curcumin-loaded flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 6
<p>Effects of different temperatures on curcumin liposomes and curcumin-loaded flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 7
<p>The effect of soy lecithin concentrations on flexible nano-liposomes (cholesterol concentration: 0.15 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 8
<p>The effect of cholesterol concentrations on flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; curcumin concentration: 0.20 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">Figure 9
<p>The effect of curcumin concentrations on curcumin-loaded flexible nano-liposomes (soy lecithin concentration: 12 mg/mL; cholesterol concentration: 0.15 mg/mL; Tween 80 concentration: 0.1 wt%).</p>
Full article ">
17 pages, 3778 KiB  
Article
High-Performance Ammonia QCM Sensor Based on SnO2 Quantum Dots/Ti3C2Tx MXene Composites at Room Temperature
by Chong Li, Ran Tao, Jinqiao Hou, Huanming Wang, Chen Fu and Jingting Luo
Nanomaterials 2024, 14(22), 1835; https://doi.org/10.3390/nano14221835 - 16 Nov 2024
Viewed by 410
Abstract
Ammonia (NH3) gas is prevalent in industrial production as a health hazardous gas. Consequently, it is essential to develop a straightforward, reliable, and stable NH3 sensor capable of operating at room temperature. This paper presents an innovative approach to modifying [...] Read more.
Ammonia (NH3) gas is prevalent in industrial production as a health hazardous gas. Consequently, it is essential to develop a straightforward, reliable, and stable NH3 sensor capable of operating at room temperature. This paper presents an innovative approach to modifying SnO2 colloidal quantum dots (CQDs) on the surface of Ti3C2Tx MXene to form a heterojunction, which introduces a significant number of adsorption sites and enhances the response of the sensor. Zero-dimensional (0D) SnO2 quantum dots and two-dimensional (2D) Ti3C2Tx MXene were prepared by solvothermal and in situ etching methods, respectively. The impact of the mass ratio between two materials on the performance was assessed. The sensor based on 12 wt% Ti3C2Tx MXene/SnO2 composites demonstrates excellent performance in terms of sensitivity and response/recovery speed. Upon exposure to 50 ppm NH3, the frequency shift in the sensor is −1140 Hz, which is 5.6 times larger than that of pure Ti3C2Tx MXene and 2.8 times higher than that of SnO2 CQDs. The response/recovery time of the sensor for 10 ppm NH3 was 36/54 s, respectively. The sensor exhibited a theoretical detection limit of 73 ppb and good repeatability. Furthermore, a stable sensing performance can be maintained after 30 days. The enhanced sensor performance can be attributed to the abundant active sites provided by the accumulation/depletion layer in the Ti3C2Tx/SnO2 heterojunction, which facilitates the adsorption of oxygen molecules. This work promotes the gas sensing application of MXenes and provides a way to improve gas sensing performance. Full article
Show Figures

Figure 1

Figure 1
<p>Schematics of the synthesis of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites and preparation of ammonia QCM sensors.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) accordion-like Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> MXene, (<b>d</b>) SnO<sub>2</sub> CQDs, and (<b>g</b>) Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites. Inset of (<b>a</b>) is SEM image of MXene with few layers. HR-TEM images of (<b>b</b>,<b>c</b>) MXene, (<b>e</b>,<b>f</b>) SnO<sub>2</sub> CQDs, and (<b>h</b>) Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites. (<b>i</b>) SAED pattern of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites. (<b>j</b>) EDS map scanning analysis of Sn, O, Ti, and C elements.</p>
Full article ">Figure 3
<p>XRD patterns of the Ti<sub>3</sub>AlC<sub>2</sub> MAX phase, accordion-like Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> MXene, pure SnO<sub>2</sub> CQDs, and Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites.</p>
Full article ">Figure 4
<p>XPS spectra of (<b>a</b>) C 1s, (<b>b</b>) O 1s, (<b>c</b>) Sn 3d, and (<b>d</b>) Ti 2p.</p>
Full article ">Figure 5
<p>Response-time curves with different NH<sub>3</sub> concentrations based on (<b>a</b>) a Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> sensor, (<b>b</b>) a SnO<sub>2</sub> sensor, (<b>c</b>) a Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> sensor. (<b>d</b>) The curves between frequency shift and gas concentration of all sensors.</p>
Full article ">Figure 6
<p>(<b>a</b>) Frequency shift of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> sensor to NH<sub>3</sub> concentrations varying from 200 ppb to 8 ppm. Response and recovery curves of (<b>b</b>) SnO<sub>2</sub>, (<b>c</b>) Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>, and (<b>d</b>) Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites.</p>
Full article ">Figure 7
<p>(<b>a</b>) Repeated response of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> sensor to 10 ppm NH<sub>3</sub>. (<b>b</b>) Selectivity of sensor based on Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites. (<b>c</b>) Response curves of Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> sensor to 10 ppm of NH<sub>3</sub> at different relative humidities. (<b>d</b>) Long-term stability of sensor based on Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites.</p>
Full article ">Figure 8
<p>Energy band diagrams of SnO<sub>2</sub> and Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> (<b>a</b>) before and (<b>b</b>) after contact. Schematic diagram of NH<sub>3</sub> sensing mechanism of sensor based on Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>/SnO<sub>2</sub> composites, (<b>c</b>) in air, and (<b>d</b>) exposure to NH<sub>3</sub>.</p>
Full article ">
15 pages, 5249 KiB  
Article
A Comprehensive Microstructure-Aware Electromigration Modeling Framework; Investigation of the Impact of Trench Dimensions in Damascene Copper Interconnects
by Ahmed Sobhi Saleh, Kristof Croes, Hajdin Ceric, Ingrid De Wolf and Houman Zahedmanesh
Nanomaterials 2024, 14(22), 1834; https://doi.org/10.3390/nano14221834 - 16 Nov 2024
Viewed by 291
Abstract
As electronic devices continue to shrink in size and increase in complexity, the current densities in interconnects drastically increase, intensifying the effects of electromigration (EM). This renders the understanding of EM crucial, due to its significant implications for device reliability and longevity. This [...] Read more.
As electronic devices continue to shrink in size and increase in complexity, the current densities in interconnects drastically increase, intensifying the effects of electromigration (EM). This renders the understanding of EM crucial, due to its significant implications for device reliability and longevity. This paper presents a comprehensive simulation framework for the investigation of EM in nano-interconnects, with a primary focus on unravelling the influential role of microstructure, by considering the impact of diffusion heterogeneity through the metal texture and interfaces. As such, the resulting atomic flux and stress distribution within nano-interconnects could be investigated. To this end, a novel approach to generate microstructures of the conductor metal is presented, whereby a predefined statistical distribution of grain sizes obtained from experimental texture analyses can be incorporated into the presented model, making the model predictive under various scales and working conditions with no need for continuous calibration. Additionally, the study advances beyond the state-of-the-art by comprehensively simulating all stages of electromigration including stress evolution, void nucleation, and void dynamics. The model was employed to study the impact of trench dimensions on the dual damascene copper texture and its impact on electromigration aging, where the model findings were corroborated by comparing them to the available experimental findings. A nearly linear increase in normalized time to nucleation was detected as the interconnect became wider with a fixed height for aspect ratios beyond 1. However, a saturation was detected with a further increase in width for lines of aspect ratios below 1, with no effective enhancement in time to nucleation. An aspect ratio of 1 seems to maximize the EM lifetime for a fixed cross-sectional area by fostering a bamboo-like structure, where about a 2-fold of increase was estimated when going from aspect ratio 2 to 1. Full article
(This article belongs to the Special Issue Mechanical and Thermal Properties of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>A comparison between the atomic flux model in (<b>A</b>) the homogenous effective diffusivity approach [<a href="#B8-nanomaterials-14-01834" class="html-bibr">8</a>] and (<b>B</b>) the presented modeling approach in this work.</p>
Full article ">Figure 2
<p>The model utilized to generate grain microstructures. (<b>A</b>) The packed circles together with their weighted Voronoi tessellation. (<b>B</b>) The created grains from the tessellation after applying the grain boundary thickness. (<b>C</b>) The grain size distribution of the generated grains (orange histogram) shown together with the lognormal grain size distribution (continuous black curve).</p>
Full article ">Figure 3
<p>An example of the meshing scheme applied to our model. Two successive zoom-ins show how the mesh size varies from triple points to the middle of a grain boundary.</p>
Full article ">Figure 4
<p>A vector plot of the atomic flux when the electron flow is from left to right. The direction of the blue arrows indicates the atomic flux direction, and their lengths indicate the flux magnitudes.</p>
Full article ">Figure 5
<p>A colormap indicating the divergence of atomic flux through the diffusion path network. Divergence was non-existent along the path and had a positive or negative value at triple points depending on the type and angle of the path.</p>
Full article ">Figure 6
<p>A colormap plot to show the stress and the void’s geometry change with time. Two color bars are used for clarity, where the top bar has a small range to illustrate the subtle differences in stress along the grain’s boundaries and cap interfaces during the initial stress evolution phase (<b>A</b>–<b>D</b>). The bottom bar has a wide range to illustrate the drastic stress changes pre- and post-void nucleation (<b>E</b>–<b>H</b>). Multimedia is available online.</p>
Full article ">Figure 7
<p>(<b>A</b>) The stress evolution along the cap interface beyond the void nucleation time. Subplot shows the stress gradient at the void surface/cap intersection point and the current crowding factor at the void surface. (<b>B</b>) The simulated electromigration induced an increase in resistance with time, showing the three stages of electromigration aging.</p>
Full article ">Figure 8
<p>The time to nucleation normalized to the minimum value for different linewidths as estimated by our model vs. experimental results. For each simulation point, the average grain diameter, d, is shown together with a snapshot of the generated microstructure. Dashed vertical lines mark different aspect ratios. The temperature and rest of the working conditions in the simulations shown in this figure follow the reported values in [<a href="#B5-nanomaterials-14-01834" class="html-bibr">5</a>].</p>
Full article ">Figure 9
<p>The full scaling study showing time to nucleation at various linewidth and heights. Values were normalized w.r.t the (80,80) point. All simulation points along with a sigmoid fit (<b>left</b>). Interpolated surface shown by a colormap (<b>right</b>).</p>
Full article ">Figure 10
<p>The “p” factor calculated for different microstructures, generated by our calibrated microstructure generation module for different aspect ratios (black points), along with a sigmoid fit to give an analytical formula (blue line).</p>
Full article ">Figure 11
<p>Time to nucleation (normalized) vs. aspect ratio at different cross-sectional areas (<b>left</b>). Time to nucleation (normalized) vs. cross-sectional area at different aspect ratios (<b>right</b>). For each figure, an inset is given to show the lines of interest on the full scaling colormap.</p>
Full article ">Figure 12
<p>The EM-induced increase in resistance above the initial value as a function of time for different cross-sectional dimensions. Time 0 was defined as the time of void nucleation for each case.</p>
Full article ">
21 pages, 2882 KiB  
Review
Gold Nanoprobes for Robust Colorimetric Detection of Nucleic Acid Sequences Related to Disease Diagnostics
by Maria Enea, Andreia Leite, Ricardo Franco and Eulália Pereira
Nanomaterials 2024, 14(22), 1833; https://doi.org/10.3390/nano14221833 - 16 Nov 2024
Viewed by 335
Abstract
Gold nanoparticles (AuNPs) are highly attractive for applications in the field of biosensing, particularly for colorimetric nucleic acid detection. Their unique optical properties, which are highly sensitive to changes in their environment, make them ideal candidates for developing simple, rapid, and cost-effective assays. [...] Read more.
Gold nanoparticles (AuNPs) are highly attractive for applications in the field of biosensing, particularly for colorimetric nucleic acid detection. Their unique optical properties, which are highly sensitive to changes in their environment, make them ideal candidates for developing simple, rapid, and cost-effective assays. When functionalized with oligonucleotides (Au-nanoprobes), they can undergo aggregation or dispersion in the presence of complementary sequences, leading to distinct color changes that serve as a visual signal for detection. Aggregation-based assays offer significant advantages over other homogeneous assays, such as fluorescence-based methods, namely, label-free protocols, rapid interactions in homogeneous solutions, and detection by the naked eye or using low-cost instruments. Despite promising results, the application of Au-nanoprobe-based colorimetric assays in complex biological matrices faces several challenges. The most significant are related to the colloidal stability and oligonucleotide functionalization of the Au-nanoprobes but also to the mode of detection. The type of functionalization method, type of spacer, the oligo–AuNPs ratio, changes in pH, temperature, or ionic strength influence the Au-nanoprobe colloidal stability and thus the performance of the assay. This review elucidates characteristics of the Au-nanoprobes that are determined for colorimetric gold nanoparticles (AuNPs)-based nucleic acid detection, and how they influence the sensitivity and specificity of the colorimetric assay. These characteristics of the assay are fundamental to developing low-cost, robust biomedical sensors that perform effectively in biological fluids. Full article
(This article belongs to the Special Issue Noble Metal-Based Nanostructures: Optical Properties and Applications)
Show Figures

Figure 1

Figure 1
<p>Timeline of AuNPs use for nucleic acid detection.</p>
Full article ">Figure 2
<p>Dependence of LSPR on spherical gold nanoparticles diameter and aggregation state.</p>
Full article ">Figure 3
<p>Colorimetric detection methods using spherical AuNPs: (Top panel) Cross-linking assay—a color change occurs as nucleic acid sequence strands specifically hybridize with complementary sequences, reducing the distance between particles, and resulting in a blue solution (positive test). In the absence of complementary sequences, the solution stays red (negative test). (Middle panel) Non-cross-linking assay—an increase in ionic strength induces AuNP aggregation, resulting in a blue solution (negative test). When complementary targets are present, the solution stays red (positive test). (Bottom panel) Colorimetric assay using unmodified AuNPs: In the absence of complementary sequences, only single-stranded DNA (ssDNA) is present, stabilizing AuNPs against salt-induced aggregation, and the solution stays red (negative result). Conversely, when hybridization occurs in the presence of a complementary sequence, double-stranded DNA (dsDNA) forms, and aggregation occurs (blue solution is a positive result). UV/vis spectra and Nanoparticle Tracking analysis (NTA) profiles are shown with blue lines corresponding to aggregated AuNPs samples and red lines to non-aggregated ones. Also indicated are the positive (green check) and negative (red cross) results for each test.</p>
Full article ">Figure 4
<p>Published successful functionalization methods of AuNPs with HS-oligos, resulting in Au-nanoprobes.</p>
Full article ">Figure 5
<p>Examples of Au nanoparticle interaction with (i) ssDNA, (ii) PolyA-ssDNA and PolyT-ssDNA, (iii) PEG-ssDNA, and (iv) thiolated-(CH2)6-ssDNA.</p>
Full article ">
20 pages, 5333 KiB  
Article
Green Synthesis of Fe2O3 Nanoparticles Using Eucalyptus globulus Leaf Extract on Pinus radiata Sawdust for Cationic Dye Adsorption
by Pablo Salgado, Eduardo Aedo and Gladys Vidal
Nanomaterials 2024, 14(22), 1832; https://doi.org/10.3390/nano14221832 - 16 Nov 2024
Viewed by 427
Abstract
The present study reports the synthesis of Fe2O3 nanoparticles on Pinus radiata sawdust (Fe2O3@PS) using a Eucalyptus globulus leaf extract. The morphology and structure of Fe2O3@PS were characterized using scanning electron microscopy [...] Read more.
The present study reports the synthesis of Fe2O3 nanoparticles on Pinus radiata sawdust (Fe2O3@PS) using a Eucalyptus globulus leaf extract. The morphology and structure of Fe2O3@PS were characterized using scanning electron microscopy (SEM), X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), and UV–Vis diffuse reflectance. The adsorption capacity of the system was evaluated by testing its ability to remove the Rhodamine B (RhB) dye. The optimization of the system was carried out using the Plackett–Burman design (PBD) and the response surface methodology (steepest ascent and the Box–Behnken design), which provided information on the main parameters affecting the adsorption process. The PBD results showed that the most important parameters for the removal of RhB using Fe2O3@PS were the removal time, the RhB concentration, and the initial pH of the system. The reusability of Fe2O3@PS under optimal conditions was tested and it was found to maintain its efficiency after five cycles of use. The efficiency and rate of RhB removal observed at pH values near 7.0 were found to be predominantly influenced by electrostatic interactions. In contrast, the analyses conducted at pH values near 8.3 exhibited reduced influence from electrostatic attractions, with π–π interactions and hydrogen bonds emerging as dominant forces. At pH values exceeding 8.3, all potential interactions between RhB and Fe2O3@PS exhibited diminished strength. This research provides valuable information on the formation of eco-friendly nanoparticles immobilized on a forest residue such as sawdust, which can effectively remove organic pollutants like RhB. This contributes to the valorization of resources and the search for solutions to water pollution. Full article
Show Figures

Figure 1

Figure 1
<p>Photographic images of (<b>a</b>) pristine PS, and (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>@PS.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) Pristine PS, (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>@PS, and (<b>c</b>) histogram of Fe<sub>2</sub>O<sub>3</sub>@PS sizes.</p>
Full article ">Figure 3
<p>DRX analysis of (<b>a</b>) pristine PS and (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>@PS.</p>
Full article ">Figure 4
<p>FTIR analyses for pristine PS and Fe<sub>2</sub>O<sub>3</sub>@PS (The black dotted line indicates a decline in signal intensity, the green dotted line represents a shift in wavenumber, and the blue dotted line denotes the disappearance of a signal).</p>
Full article ">Figure 5
<p>Absorption spectra of Fe<sub>2</sub>O<sub>3</sub>@PS and PS.</p>
Full article ">Figure 6
<p>Pareto chart showing the standardized effects of variables on RhB adsorption.</p>
Full article ">Figure 7
<p>Residual diagnostics of the quadratic model. (<b>a</b>) normal probability plot of residuals, (<b>b</b>) residuals against the predicted values of the model, (<b>c</b>) actual and predicted values of RhB adsorption (%) based on BBD.</p>
Full article ">Figure 8
<p>Effect of (<b>a</b>) time, (<b>b</b>) RhB concentration, and (<b>c</b>) pH on the adsorption efficiency of RhB.</p>
Full article ">Figure 9
<p>Determination of pHpzc for Fe<sub>2</sub>O<sub>3</sub>@PS using pH drift method.</p>
Full article ">Figure 10
<p>Three-dimensional response surface plots representing the modeled RhB adsorption (%) as a function of (<b>a</b>) time and RhB concentration, (<b>b</b>) time and pH, and (<b>c</b>) RhB concentration and pH.</p>
Full article ">Figure 11
<p>Reusability study of RhB adsorption using Fe<sub>2</sub>O<sub>3</sub>@PS by (<b>a</b>) conditions 1, and (<b>b</b>) conditions 2. The data obtained are presented as mean ± standard deviation.</p>
Full article ">Figure 12
<p>(<b>a</b>) RhB adsorption efficiency by Fe<sub>2</sub>O<sub>3</sub>@PS at different time intervals. (<b>b</b>) Calculated RhB adsorption rate constant (k<sub>app</sub>) for Fe<sub>2</sub>O<sub>3</sub>@PS (inset: legend represents k<sub>app</sub> and r<sup>2</sup> of a pseudo-first order model).</p>
Full article ">Figure 13
<p>Possible interactions in the RhB adsorption process using Fe<sub>2</sub>O<sub>3</sub>@PS under optimal conditions.</p>
Full article ">
13 pages, 7273 KiB  
Article
Catalytic Methane Decomposition on In Situ Reduced FeCo Alloy Catalysts Derived from Layered Double Hydroxides
by Dianfeng Cao, Yuwen Li, Chao Lv, Yongtao An, Jiangfeng Song, Mingcan Li and Xin Zhang
Nanomaterials 2024, 14(22), 1831; https://doi.org/10.3390/nano14221831 - 15 Nov 2024
Viewed by 284
Abstract
Catalytic methane decomposition (CMD) reaction is considered a promising process for converting greenhouse gas CH4 into hydrogen and high-value-added carbon materials. In this work, a series of Al2O3-supported FeCo alloy catalysts were successfully prepared in the CMD process. [...] Read more.
Catalytic methane decomposition (CMD) reaction is considered a promising process for converting greenhouse gas CH4 into hydrogen and high-value-added carbon materials. In this work, a series of Al2O3-supported FeCo alloy catalysts were successfully prepared in the CMD process. Compared to the pre-reduced catalysts, the in situ reduced FeCo alloy catalysts showed higher methane conversion rates, with the highest reaching 83% at 700 °C, due to the finer active nanoparticle size and greater exposure of active site. Furthermore, the time-on-stream tests demonstrated that the catalytic activity of in situ reduced FeCo alloy catalysts could remain above 92.3% of the highest catalytic activity after 10 h. In addition, TEM analyses of the carbon products from the CMD in situ reduced catalysts revealed the production of carbon nanofibers and nanotubes several microns in length after the reaction. This indicates that the in situ reduced FeCo alloy catalysts more effectively promoted the growth of carbon nanofibers. These results could provide a viable strategy for future methane decomposition development aimed at producing hydrogen and high-value carbon. Full article
(This article belongs to the Special Issue Nanomaterials for Sustainable Green Energy)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of FeCoAl-LDH-<span class="html-italic">x</span> and (<b>b</b>) XRD patterns of FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>.</p>
Full article ">Figure 2
<p>TEM images of (<b>a</b>–<b>c</b>) FeCoAl-LDH-<span class="html-italic">x</span> (<span class="html-italic">x</span> = 1–3). (<b>d</b>–<b>f</b>) TEM images of FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span> (<span class="html-italic">x</span> = 1–3). (<b>g</b>–<b>i</b>) Statistical distribution graphs of the particle size of FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span> (<span class="html-italic">x</span> = 1–3).</p>
Full article ">Figure 3
<p>(<b>a</b>) TG analysis of the FeCoAl-LDH-<span class="html-italic">x</span>, (<b>b</b>) the differential curves of TG about FeCoAl-LDH-<span class="html-italic">x</span>, (<b>c</b>) TG analysis of the FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>, and (<b>d</b>) the differential curves of TG about FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>.</p>
Full article ">Figure 4
<p>The relationship between methane conversion and temperature in the TPSR test at 5 °C/min to 800 °C for (<b>a</b>) FeCoAl-LDH-<span class="html-italic">x</span> and (<b>b</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>, and the relationship between methane conversion and temperature in the TOS test at 5 °C/min to 700 °C for insulation of (<b>c</b>) FeCoAl-LDH-<span class="html-italic">x</span> and (<b>d</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>.</p>
Full article ">Figure 5
<p>XRD patterns of methane decomposition reaction products catalyzed by (<b>a</b>) FeCoAl-LDH-<span class="html-italic">x</span> and (<b>b</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span>.</p>
Full article ">Figure 6
<p>TEM images of carbon products created by (<b>a</b>,<b>b</b>) FeCoAl-LDH-1, (<b>c</b>,<b>d</b>) FeCoAl-LDH-2, (<b>e</b>,<b>f</b>) and FeCoAl-LDH-3.</p>
Full article ">Figure 7
<p>TEM images of TG products created by (<b>a</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-1, (<b>b</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-2, and (<b>c</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-3.</p>
Full article ">Figure 8
<p>Raman spectra of (<b>a</b>) spent FeCoAl-LDHs-<span class="html-italic">x</span> (<span class="html-italic">x</span> = 1–3) and (<b>b</b>) FeCo/Al<sub>2</sub>O<sub>3</sub>-<span class="html-italic">x</span> (<span class="html-italic">x</span> = 1–3) catalysts after TG.</p>
Full article ">
14 pages, 7597 KiB  
Article
Magnetic Field/Ultrasound-Responsive Fe3O4 Microbubbles for Targeted Mechanical/Catalytic Removal of Bacterial Biofilms
by Liang Lu, Yuan Liu, Xiaolong Chen, Fengjiao Xu, Qi Zhang, Zhaowei Yin and Lihui Yuwen
Nanomaterials 2024, 14(22), 1830; https://doi.org/10.3390/nano14221830 - 15 Nov 2024
Viewed by 336
Abstract
Conventional antibiotics are limited by drug resistance, poor penetration, and inadequate targeting in the treatment of bacterial biofilm-associated infections. Microbubble-based ultrasound (US)-responsive drug delivery systems can disrupt biofilm structures and enhance antibiotic penetration through cavitation effects. However, currently developed US-responsive microbubbles still depend [...] Read more.
Conventional antibiotics are limited by drug resistance, poor penetration, and inadequate targeting in the treatment of bacterial biofilm-associated infections. Microbubble-based ultrasound (US)-responsive drug delivery systems can disrupt biofilm structures and enhance antibiotic penetration through cavitation effects. However, currently developed US-responsive microbubbles still depend on antibiotics and lack targeting capability. In this work, magnetic field/ultrasound (MF/US)-responsive Fe3O4 microbubbles (FMB) were constructed based on Fe3O4 nanoparticles (NPs) with superparamagnetic and peroxidase-like catalytic properties. In vitro experiments demonstrated that FMB can be targeted to methicillin-resistant Staphylococcus aureus (MRSA) biofilms by the direction of MF. Upon US irradiation, FMB collapse due to inertial cavitation and generate mechanical forces to disrupt the structure of MRSA biofilms and releases Fe3O4 NPs, which catalyze the generation of reactive oxygen species (ROS) from H2O2 in the biofilm microenvironment and kill the bacteria within the biofilm. In a mouse biofilm infection model, FMB efficiently destroyed MRSA biofilms grown in subcutaneous catheters with the MF and US. Magnetic-targeted mechanical/catalytic therapy based on FMB provides a promising strategy for effectively combating bacterial biofilm infection. Full article
(This article belongs to the Special Issue Stimuli-Responsive Nanomaterials for Imaging and Therapy)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of magnetic field/ultrasound (MF/US)-responsive Fe<sub>3</sub>O<sub>4</sub> microbubbles (FMB) for bacterial biofilm removal. (<b>a</b>) Mechanism of action of FMB for MF-targeted mechanical/catalytic removal of bacterial biofilms. (<b>b</b>) FMB target the MRSA biofilm of mouse subcutaneous catheter under the guidance of MF, destroy the biofilm structure by ultrasound cavitation effect, and catalyze the production of ROS from H<sub>2</sub>O<sub>2</sub> to kill the bacteria in the biofilm.</p>
Full article ">Figure 2
<p>Preparation and characterization of FMB. (<b>a</b>) Schematic of the preparation process of FMB. (<b>b</b>) Bright-field microphotograph of FMB. (<b>c</b>) Size distribution histogram of FMB with a statistical number greater than 200. (<b>d</b>) Content of Fe in different volumes of FMB aqueous dispersion. (<b>e</b>) Scanning electron microscopy (SEM) images and elemental mapping images of FMB. (<b>f</b>) X-ray diffraction (XRD) spectra of Fe<sub>3</sub>O<sub>4</sub> NPs, FMB, and standard powder diffraction pattern of Fe<sub>3</sub>O<sub>4</sub> (PDF#04-006-0424).</p>
Full article ">Figure 3
<p>MF/US-responsive properties of FMB. (<b>a</b>) Photographs of FMB before and after ultrasound irradiation. (<b>b</b>) Photographs of FMB before and after the action of MF. (<b>c</b>) Vibrating sample magnetometer (VSM) spectra of FMB. (<b>d</b>) Photographs of FMB migrating to the target position in the catheter under the action of a permanent magnet. t1, t2, t3, and t4 represent different time points. The red arrow indicates the moving direction of FMB.</p>
Full article ">Figure 4
<p>Catalytic properties of FMB. (<b>a</b>) Photographs of FMB catalyzing TMB in a Petri dish containing 300 mM H<sub>2</sub>O<sub>2</sub>. P1 and P2 represent the locations of FMB at different time points. The yellow arrow indicates the moving path of FMB. (<b>b</b>) Ultraviolet-visible-near-infrared (UV-vis-NIR) absorption spectra of FMB dispersions after reaction with TMB under different conditions; insets from left to right are photographs of TMB + H<sub>2</sub>O<sub>2</sub>, TMB + H<sub>2</sub>O<sub>2</sub> + FMB, and TMB + H<sub>2</sub>O<sub>2</sub> + FMB + US groups.</p>
Full article ">Figure 5
<p>MRSA biofilm disruption by FMB under MF and US. (<b>a</b>) Optical photographs, (<b>b</b>) microphotographs, and (<b>c</b>) relative biofilm biomass of MRSA biofilms after crystal violet staining with different treatments.</p>
Full article ">Figure 6
<p>MRSA biofilm disruption by FMB under MF and US at different concentrations of H<sub>2</sub>O<sub>2</sub>. (<b>a</b>) Optical photographs, (<b>b</b>) microphotographs, and (<b>c</b>) relative biofilm biomass of MRSA biofilms grown in 96-well plates with crystal violet staining after different treatments.</p>
Full article ">Figure 7
<p>Fluorescence imaging of MRSA biofilms. (<b>a</b>) Three-dimensional confocal laser scanning microscopy (3D CLSM) photographs of MRSA biofilms stained by Calcein–AM after different treatments. (<b>b</b>) Thickness of MRSA biofilms after different treatments. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Antibacterial effect of FMB against MRSA biofilms. (<b>a</b>) Photographs of MRSA colonies on agar plates and (<b>b</b>) the number of viable bacteria in MRSA biofilms after different treatments. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>In vivo catheter MRSA biofilm clearance by FMB. (<b>a</b>) Schematic of the treatment of catheter biofilms in mice. (<b>b</b>) Photographs of crystal violet-stained and (<b>c</b>) relative biofilm biomass of catheters after different treatments. (<b>d</b>) MRSA colonies on agar plates and (<b>e</b>) number of viable bacteria within MRSA biofilm. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
3 pages, 2535 KiB  
Correction
Correction: Arbi et al. Polypyrrole-Assisted Ag Doping Strategy to Boost Co(OH)2 Nanosheets on Ni Foam as a Novel Electrode for High-Performance Hybrid Supercapacitors. Nanomaterials 2022, 12, 3982
by Hammad Mueen Arbi, Anuja A. Yadav, Yedluri Anil Kumar, Md Moniruzzaman, Salem Alzahmi and Ihab M. Obaidat
Nanomaterials 2024, 14(22), 1829; https://doi.org/10.3390/nano14221829 - 15 Nov 2024
Viewed by 182
Abstract
In the original publication [...] Full article
(This article belongs to the Special Issue Nanostructured Materials for Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the synthesis of Ag-doped@Co(OH)<sub>2</sub>@polypyrrole NSs.</p>
Full article ">Figure 2
<p>SEM images of the prepared Co(OH)<sub>2</sub> nanoparticles (<b>a</b>–<b>c</b>), Ag-doped@Co(OH)<sub>2</sub> nanoparticles (<b>d</b>–<b>f</b>), and Ag-doped@Co(OH)<sub>2</sub>@polypyrrole NSs (<b>g</b>–<b>i</b>).</p>
Full article ">Figure 4
<p>The XRD patterns for the electrode materials (<b>a</b>); the XPS survey spectrum (<b>b</b>) and the XPS spectra of Co 2p (<b>c</b>), O 1 s (<b>d</b>), and Ag 3d (<b>e</b>) of the Ag-doped@Co(OH)<sub>2</sub>@polypyrrole NSs.</p>
Full article ">
34 pages, 4568 KiB  
Review
Nanothermodynamics: There’s Plenty of Room on the Inside
by Ralph V. Chamberlin and Stuart M. Lindsay
Nanomaterials 2024, 14(22), 1828; https://doi.org/10.3390/nano14221828 - 15 Nov 2024
Viewed by 385
Abstract
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical [...] Read more.
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical scaling near ferromagnetic transitions, thermal and dynamic behavior near liquid–glass transitions, and the 1/f-like noise in metal films and qubits. A key feature in several models is to allow separate time steps for distinct conservation laws: one type of step conserves energy and the other conserves momentum (e.g., dipole alignment). This “orthogonal dynamics” explains how the relaxation of a single parameter can exhibit multiple responses such as primary, secondary, and microscopic peaks in the dielectric loss of supercooled liquids, and the crossover in thermal fluctuations from Johnson–Nyquist (white) noise at high frequencies to 1/f-like noise at low frequencies. Nanothermodynamics also provides new insight into three basic questions. First, it gives a novel solution to Gibbs’ paradox for the entropy of the semi-classical ideal gas. Second, it yields the stable equilibrium of Ising’s original model for finite-sized chains of interacting binary degrees of freedom (“spins”). Third, it confronts Loschmidt’s paradox for the arrow of time, showing that an intrinsically irreversible step is required for maximum entropy and the second law of thermodynamics, not only in the thermodynamic limit but also in systems as small as N=2 particles. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Finite-size thermal effects. Inset gives Hill’s fundamental equation of small-system thermodynamics, with a simple (three-energy-level) diagram for each term (adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>]). The first three terms on the right side (black) give the standard ways to increase the total internal energy of a system: add heat (<math display="inline"><semantics> <mrow> <mi>T</mi> <mi>d</mi> <msub> <mrow> <mi>S</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), do work on the system (<math display="inline"><semantics> <mrow> <mo>−</mo> <mi>P</mi> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), or add particles (<math display="inline"><semantics> <mrow> <mi>μ</mi> <mi>d</mi> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>). The fourth term (red) contains finite-size effects (surface states, length-scale terms, fluctuations, etc.) that change the width of the levels when the number of subdivisions changes if the subdivision potential is nonzero (<math display="inline"><semantics> <mrow> <mo>ℇ</mo> <mo>≠</mo> <mn>0</mn> </mrow> </semantics></math>). The main figure shows how free energy might change with the number of subdivisions, from <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in the thermodynamic limit of no subdivisions (<math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) to <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">n</mi> </mrow> </semantics></math> in the nanothermodynamic limit for stable equilibrium of subsystems inside bulk samples (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">ℇ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Schematic representation of various multiplicities. A canonical system (<b>top</b>) has two indistinguishable particles that may be on the left side (L), right side (R), or opposite sides. There is only one way to subdivide this system into canonical subsystems (<b>middle</b>), but there are many ways to subdivide it into nanocanonical subsystems (<b>bottom</b>). Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>Sketch showing two solutions to Gibbs’ paradox for combining two types of particles: X’s (blue) and O’s (red). (<b>A</b>–<b>C</b>) Canonical ensemble, where all particles of the same type are indistinguishable over all distances. (<b>D</b>–<b>F</b>) Nanocanonical ensemble, comprised of nanoscale subsystems, where similar particles can be distinguished by their location when in different subsystems. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 4
<p>Sketch showing a stable solution of the 1D Ising model at a given <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math>. Ten spins are in the chain. Each spin may be up or down. Each interaction between neighboring spins may be low energy (<math display="inline"><semantics> <mrow> <mo>●</mo> </mrow> </semantics></math>), high energy (<b>X</b>), or a no-energy “break” (<b>O</b>) in the interaction.</p>
Full article ">Figure 5
<p>(<b>D</b>) Temperature dependence of the effective scaling exponent from data (symbols) and models (lines) sketched in (<b>A</b>–<b>C</b>). Each red box encloses a separate set of spins that can be treated using mean-field theory. (<b>A</b>) Standard mean-field theory yields <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (dotted line in (<b>D</b>)). (<b>B</b>) Simulations of the standard Ising model yield a monotonic increase in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> with decreasing <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> (dashed line in (<b>D</b>)). (<b>C</b>) The mean-field cluster model yields non-monotonic behavior in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> (solid lines in (<b>D</b>)), similar to measurements on EuO (circles) and Gd (squares). Difficulty in determining <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math> yields uncertainty as <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>→</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>, but not for <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> <mo>[</mo> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>T</mi> <mo>−</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>)</mo> <mo>/</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>]</mo> </mrow> </mrow> <mo>&gt;</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> where <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> of the standard Ising model shows only gradual and monotonic behavior, unlike the measurements. Adapted from [<a href="#B26-nanomaterials-14-01828" class="html-bibr">26</a>].</p>
Full article ">Figure 6
<p>Log-log plot of frequency-dependent loss from the orthogonal Ising model. The loss is deduced from the power spectral density (PSD) using the fluctuation-dissipation theorem. The frequency is normalized by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> to put the microscopic peak at <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>f</mi> <mo>/</mo> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> <mo>~</mo> <mn>12</mn> </mrow> </semantics></math>. Simulations are made on subsystems of two sizes, each at two temperatures, as given in the legends. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 7
<p>Primary response time of glycerol. Abscissa is inverse temperature, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>/</mo> <mi>T</mi> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> is the mean-field critical temperature. The ordinate in (<b>A</b>) is <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, and in (<b>B</b>) it comes from a type of Stickel plot [<a href="#B84-nanomaterials-14-01828" class="html-bibr">84</a>] utilizing finite differences of <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, which removes the prefactor and linearize the VFT2 function. Symbols are from measurements (Stickel [<a href="#B85-nanomaterials-14-01828" class="html-bibr">85</a>]). Various lines are from the VFT2 function Equation (6) (black), VFT function (red), and MYEGA function (blue) [<a href="#B86-nanomaterials-14-01828" class="html-bibr">86</a>]. The inset is a sketch of a simple free-energy diagram, containing two minima separated by a barrier. Primary response in the orthogonal Ising model involves fluctuations in energy that open pathways between the minima. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 8
<p>1/<span class="html-italic">f</span>-like noise from maintaining maximum entropy during equilibrium fluctuations. (<b>A</b>–<b>E</b>) Sketch of all distinct configurations of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> spins, arranged in order of decreasing alignment from <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> (<b>top</b>) to <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> (<b>bottom</b>). The multiplicity for the alignment entropy of the subsystem comes from the number of configurations in each box. (<b>F</b>) Temperature-dependent exponent for noise that varies as a function of frequency, <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>S</mi> <mi>D</mi> <mo>(</mo> <mi>f</mi> <mo>)</mo> <mo>∝</mo> <mn>1</mn> <mo>/</mo> <msup> <mrow> <mi>f</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msup> </mrow> </semantics></math>, with the abscissa normalized by <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>. Solid symbols (color) are from measurements [<a href="#B88-nanomaterials-14-01828" class="html-bibr">88</a>] of noise in thin films for the metals given in the legend. Open symbols (black) are from simulations of a 3D Ising subsystems having <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>27</mn> </mrow> </semantics></math> spins with dynamics utilizing a local bath to maintain maximum entropy during fluctuations in alignment. Solid line is the best linear fit to the simulations, weighted by the inverse variance of each point. Dashed line is from a random fluctuation model [<a href="#B89-nanomaterials-14-01828" class="html-bibr">89</a>]. Adapted from [<a href="#B60-nanomaterials-14-01828" class="html-bibr">60</a>].</p>
Full article ">Figure 9
<p>Influence of energy on the amplitude of alignment fluctuations via orthogonal dynamics. (<b>A</b>–<b>E</b>) Configurations of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> interactions, arranged in order of decreasing energy. (<b>F</b>) Simulation of energy (<math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> </mrow> </semantics></math>, red) and magnetization (<math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math>, black) as a function of time for the 1D Ising model containing <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> interactions, with a local bath to maintain maximum entropy. Note how the amplitude of fluctuations in <math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math> tends to be slightly larger when <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> <mo>&lt;</mo> <mo>−</mo> <mn>0.3</mn> </mrow> </semantics></math>. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 10
<p>Noise power spectral densities from simulations (lines) and measurements (symbols). Solid lines are from fluctuations in alignment of 1D chains of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> Ising spins using orthogonal dynamics while maintaining maximum entropy. Note that <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> (blue) is a small enough subsystem to show separate Lorentzians in a 1/<span class="html-italic">f</span>-like spectrum, while <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> (red) is large enough to show a crossover from white noise at high frequencies (dotted) to 1/<span class="html-italic">f</span>-like noise at low frequencies with an exponent of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.92</mn> </mrow> </semantics></math> (dashed). Symbols are from measurements of flux noise (solid) and tunnel-coupling noise (open) in a qubit [<a href="#B92-nanomaterials-14-01828" class="html-bibr">92</a>]. Each set of measurements has been shifted in amplitude and frequency to match the simulations. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 11
<p>Time dependence of entropies per particle (<b>A</b>–<b>E</b>) and inverse effective temperatures (<b>F</b>). Simulations utilize a Creutz-like model of 1D Ising-like spins coupled to a <span class="html-italic">ke</span> bath of Einstein oscillators. Top three left-side graphs show the time-dependence of <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>/</mo> <mo>(</mo> <mi>N</mi> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> for the spins (<b>C</b>), <span class="html-italic">ke</span> bath (<b>B</b>), and their sum (<b>A</b>) in a large system, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>. Symbols come from first averaging 10,000 sweeps, then averaging three separate simulations of each type, with error bars visible if larger than the symbol size. A simulation with irreversible dynamics (red circles) precedes every simulation with reversible dynamics (black squares). Thus, the total entropy always decreases when the dynamics becomes reversible, as indicated by the orange arrow in (<b>A</b>). Furthermore, when the rate of break-change attempts is reduced to 1/10 the rate of spin-change attempts (middle third of every simulation), reversible simulations have an entropy that depends on the dynamics. Right-side graphs show the total entropies, as in (<b>A</b>) but without time-averaging, over a greatly expanded time scale. Here the differences between reversible (black) and irreversible (red) behavior are clearly visible at the start (<b>D</b>) and end (<b>E</b>) of the simulations. The inset shows corresponding differences in the power-spectral densities of the simulations. Symbols in (<b>F</b>) give the logarithm of the ratio of probabilities of neighboring energy levels in the <span class="html-italic">ke</span> bath, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> <mo>/</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> <mo>+</mo> <mn>1</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (squares), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (circles), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> (up triangles), and <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (down triangles). These values are proportional to the difference in inverse effective temperature of the adjacent levels. A single temperature applies only to irreversible dynamics in the thermodynamic limit (red), not for reversible dynamics in this limit (black) nor for irreversible dynamics of small subsystems, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>128</mn> </mrow> </semantics></math> (green). Adapted from [<a href="#B16-nanomaterials-14-01828" class="html-bibr">16</a>].</p>
Full article ">Figure 12
<p>Fluctuations in potential energy from MD simulations of Lennard–Jones crystals. Main figure shows normalized <span class="html-italic">pe</span> fluctuations for blocks of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>32</mn> </mrow> </semantics></math> atoms in a system of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>442,368</mn> </mrow> </semantics></math> atoms as a function interaction cutoff radius, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>, at three temperatures given in the legend. Note that the data (open symbols) tend to be relatively constant (independent of <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>) when interactions are robustly harmonic, having interaction between nearest-neighbor atoms only, <math display="inline"><semantics> <mrow> <mn>1.12</mn> <mo>≈</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≤</mo> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>≤</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>4</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≈</mo> <mn>1.59</mn> </mrow> </semantics></math>. Insets show the time dependence of energy autocorrelations in blocks (black squares) and energy correlations between nearest-neighbor blocks (red circles). Simulations are made at <math display="inline"><semantics> <mrow> <mi>k</mi> <mi>T</mi> <mo>/</mo> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mn>0.0005</mn> </mrow> </semantics></math> for blocks containing a single unit cell of the crystal, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>. The lower inset shows that neighboring blocks are positively correlated when all atoms have robustly harmonic interactions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>), while the upper inset (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>2.0</mn> </mrow> </semantics></math>) shows that neighboring blocks are anticorrelated when interactions include second-neighbor atoms that are anharmonic. Adapted from [<a href="#B15-nanomaterials-14-01828" class="html-bibr">15</a>] with permission from Elsevier.</p>
Full article ">
15 pages, 4260 KiB  
Article
Microwave-Assisted Synthesis of N, S Co-Doped Carbon Quantum Dots for Fluorescent Sensing of Fe(III) and Hydroquinone in Water and Cell Imaging
by Zhaochuan Yu, Chao Deng, Wenhui Ma, Yuqian Liu, Chao Liu, Tingwei Zhang and Huining Xiao
Nanomaterials 2024, 14(22), 1827; https://doi.org/10.3390/nano14221827 - 14 Nov 2024
Viewed by 503
Abstract
The detection of heavy metal ions and organic pollutants from water sources remains critical challenges due to their detrimental effects on human health and the environment. Herein, a nitrogen and sulfur co-doped carbon quantum dot (NS-CQDs) fluorescent sensor was developed using a microwave-assisted [...] Read more.
The detection of heavy metal ions and organic pollutants from water sources remains critical challenges due to their detrimental effects on human health and the environment. Herein, a nitrogen and sulfur co-doped carbon quantum dot (NS-CQDs) fluorescent sensor was developed using a microwave-assisted carbonization method for the detection of Fe3+ ions and hydroquinone (HQ) in aqueous solutions. NS-CQDs exhibit excellent optical properties, enabling sensitive detection of Fe3+ and HQ, with detection limits as low as 3.40 and 0.96 μM. Notably, with the alternating introduction of Fe3+ and HQ, NS-CQDs exhibit significant fluorescence (FL) quenching and recovery properties. Based on this property, a reliable “on-off-on” detection mechanism was established, enabling continuous and reversible detection of Fe3+ and HQ. Furthermore, the low cytotoxicity of NS-CQDs was confirmed through successful imaging of HeLa cells, indicating their potential for real-time intracellular detection of Fe3+ and HQ. This work not only provides a green and rapid synthesis strategy for CQDs but also highlights their versatility as fluorescent probes for environmental monitoring and bioimaging applications. Full article
(This article belongs to the Special Issue Nanomaterials in Electrochemical Electrode and Electrochemical Sensor)
Show Figures

Figure 1

Figure 1
<p>Synthesis and characterization of NS-CQDs. (<b>a</b>) HRTEM image of NS-CQDs (the inset shows its lattice fringes). (<b>b</b>) Diameter distribution analysis, (<b>c</b>) XRD, (<b>d</b>) AFM image, (<b>e</b>) height distribution analysis, and (<b>f</b>) Raman spectroscopy of NS-CQDs.</p>
Full article ">Figure 2
<p>Chemical composition analysis of NS-CQDs. (<b>a</b>) FTIR and (<b>b</b>) XPS spectra of NS-CQDs. (<b>c</b>–<b>f</b>) High-resolution XPS of C 1s, N 1s, S 2p, and O 1s, respectively.</p>
Full article ">Figure 3
<p>Evaluation of FL properties of NS-MQDs. (<b>a</b>) UV-vis absorption spectra and FL excitation (Ex) and FL emission (Em) spectra of NS-MQDs in aqueous solution. (<b>b</b>) FL emission spectra of NS-MQDs at different excitation wavelengths (320–480 nm).</p>
Full article ">Figure 4
<p>“On-off-on” detection performance of NS-CQDs for Fe<sup>3+</sup> and HQ. (<b>a</b>) The concentration-dependent emission spectra of NS-CQDs at different concentrations of Fe<sup>3+</sup> (0–833.3 μM). (<b>b</b>) Stern-Volmer plot of F/F<sub>0</sub> versus Fe<sup>3+</sup> concentration for NS-CQDs. (<b>c</b>) FL intensity change profiles of NS-CQDs in the presence of potential competing ions (green) and Fe<sup>3+</sup> upon addition of competing anions (3-fold excess, purple) in DI water. (<b>d</b>) The concentration-dependent emission spectra of (NS-CQDs+Fe<sup>3+</sup>) for different concentrations of HQ (0–1333.3 μM). (<b>e</b>) Stern-Volmer plot of F/F<sub>0</sub> versus HQ concentration for (NS-CQDs+Fe<sup>3+</sup>). (<b>f</b>) FL intensity change profiles of NS-CQDs in the presence of potential competing ions (green) and HQ upon addition of competing anions (3-fold excess, purple) in DI water.</p>
Full article ">Figure 5
<p>Fluorescence “on-off-on” principle of NS-CQDs. (<b>a</b>) UV-vis absorption spectrum and (<b>b</b>) FL decay curves of NS-CQDs before and after the addition of Fe<sup>3+</sup>. (<b>c</b>) FL spectrum of NS-CQDs, NS-CQDs+HQ, NS-CQDs+Fe<sup>3+</sup>, NS-CQDs Fe<sup>2+</sup>, and NS-CQDs+Fe<sup>3+</sup>+HQ. (<b>d</b>) Schematic of the proposed “on-off-on” mechanism for the detection of Fe<sup>3+</sup> and HQ using NS-CQDs.</p>
Full article ">Figure 6
<p>Confocal FL images of living HeLa cells. The images after incubating NS-CQDs at 37 °C for 4 h are as follows: (<b>a</b>) bright field, (<b>d</b>) FL, and (<b>g</b>) merged image. The images of HeLa cells stained with NS-CQDs after treatment with Fe<sup>3+</sup> (300 μM) for 4 h are: (<b>b</b>) bright field, (<b>e</b>) FL, and (<b>h</b>) merged image. The images of HeLa cells stained with NS-CQDs after treatment with Fe<sup>3+</sup> (300 μM) followed by HQ (300 μM) for 4 h are: (<b>c</b>) bright field, (<b>f</b>) FL, and (<b>i</b>) merged image.</p>
Full article ">Scheme 1
<p>Schematic of the fabrication of co-doped NS-CQDs and its applications.</p>
Full article ">
11 pages, 4207 KiB  
Article
Respiration Monitoring Using Humidity Sensor Based on  Hydrothermally Synthesized Two-Dimensional MoS2
by Gwangsik Hong, Mi Eun Kim, Jun Sik Lee, Ja-Yeon Kim and Min-Ki Kwon
Nanomaterials 2024, 14(22), 1826; https://doi.org/10.3390/nano14221826 - 14 Nov 2024
Viewed by 494
Abstract
Breathing is the process of exchanging gases between the human body and the surrounding environment. It plays a vital role in maintaining human health, sustaining life, and supporting various bodily functions. Unfortunately, current methods for monitoring respiration are impractical for medical applications because [...] Read more.
Breathing is the process of exchanging gases between the human body and the surrounding environment. It plays a vital role in maintaining human health, sustaining life, and supporting various bodily functions. Unfortunately, current methods for monitoring respiration are impractical for medical applications because of their high costs and need for bulky equipment. When measuring changes in moisture during respiration, we observed a slow response time for 2D nanomaterial-based resistance measurement methods used in respiration sensors. Through thermal annealing, the crystal structure of MoS2 is transformed from 1T@2H to 2H, allowing the measurement of respiration at more than 30 cycles per minute and enabling analysis of the response. This study highlights the potential of two-dimensional nanomaterials for the development of low-cost and highly sensitive humidity and respiration sensors for various applications. Full article
(This article belongs to the Special Issue 2D Materials for Advanced Sensors: Fabrication and Applications)
Show Figures

Figure 1

Figure 1
<p>The fabrication process steps for the MaS<sub>2</sub>-based respiration sensor.</p>
Full article ">Figure 2
<p>Optical microscope images at different annealing temperatures: (<b>a</b>) before annealing, (<b>b</b>) at 400 °C, (<b>c</b>) at 700 °C; (<b>d</b>) Raman spectrums of MOS<sub>2</sub> without and with annealing (400 and 700 °C) and; (<b>e</b>) peak frequency difference between A<sub>1g</sub> and E<sup>1</sup><sub>2g</sub> peaks.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD spectra and S 2p and Mo 3d XPS spectra of MoS<sub>2</sub> (<b>b</b>) with and (<b>c</b>) without thermal annealing.</p>
Full article ">Figure 4
<p>TEM images of MoS<sub>2</sub> annealed at 700 °C: (<b>a</b>) low magnification and (<b>b</b>) high-resolution TEM image.</p>
Full article ">Figure 5
<p>Current response to humidity change for resistive-type, MoS<sub>2</sub>-based respiration sensors (<b>a</b>) without and (<b>b</b>) with thermal annealing.</p>
Full article ">Figure 6
<p>Current response of MoS<sub>2</sub>-based sensors depending on distance between nose and sensor: (<b>a</b>) without thermal annealing and (<b>b</b>) with thermal annealing; and current response to normal and fast breathing with MoS<sub>2</sub>: (<b>c</b>) without thermal annealing and (<b>d</b>) with thermal annealing.</p>
Full article ">Figure 7
<p>Current response of single breathing with sensor using MoS<sub>2</sub> (<b>a</b>) without and (<b>b</b>) with thermal annealing.</p>
Full article ">Figure 8
<p>(<b>a</b>) Mice without cancer cell injection; (<b>b</b>) mice injected with cancer cells; and (<b>c</b>) mouse set up for measurement of respiration response. The red dot ring in (<b>b</b>) shows the location and appearance of cancer cells.</p>
Full article ">Figure 9
<p>Respiratory responses of healthy mice (<b>a</b>,<b>b</b>) and cancer-bearing mice (<b>c</b>,<b>d</b>): (<b>a</b>) week 3, (<b>b</b>) zoom in data of (<b>a</b>), week 3, (<b>c</b>) week 0, (<b>d</b>) week 3.</p>
Full article ">
2 pages, 170 KiB  
Editorial
Semiconductor Quantum Dots: Synthesis, Properties and Applications
by Donghai Feng, Guofeng Zhang and Yang Li
Nanomaterials 2024, 14(22), 1825; https://doi.org/10.3390/nano14221825 - 14 Nov 2024
Viewed by 510
Abstract
Semiconductor nanoparticles of sizes smaller than exciton Bohr diameters undergo quantum confinement and are called quantum dots (QDs), which exhibit size-dependent physicochemical properties [...] Full article
(This article belongs to the Special Issue Semiconductor Quantum Dots: Synthesis, Properties and Applications)
18 pages, 5175 KiB  
Article
Co-Activating Lattice Oxygen of TiO2-NT and SnO2 Nanoparticles on Superhydrophilic Graphite Felt for Boosting Electrocatalytic Oxidation of Glyphosate
by Wenyan He, Sheng Bai, Kaijie Ye, Siyan Xu, Yinuo Dan, Moli Chen and Kuo Fang
Nanomaterials 2024, 14(22), 1824; https://doi.org/10.3390/nano14221824 - 14 Nov 2024
Viewed by 267
Abstract
Glyphosate (GH) wastewater potentially poses hazards to human health and the aquatic environment, due to its persistence and toxicity. A highly superhydrophilic and stable graphite felt (GF)/polydopamine (PDA)/titanium dioxide nanotubes (TiO2-NT)/SnO2/Ru anode was fabricated and characterized for the degradation [...] Read more.
Glyphosate (GH) wastewater potentially poses hazards to human health and the aquatic environment, due to its persistence and toxicity. A highly superhydrophilic and stable graphite felt (GF)/polydopamine (PDA)/titanium dioxide nanotubes (TiO2-NT)/SnO2/Ru anode was fabricated and characterized for the degradation of glyphosate wastewater. Compared to control anodes, the GF/PDA/TiO2-NT/SnO2/Ru anode exhibited the highest removal efficiency (near to 100%) and a yield of phosphate ions of 76.51%, with the lowest energy consumption (0.088 Wh/L) for degrading 0.59 mM glyphosate (GH) at 7 mA/cm2 in 30 min. The exceptional activity of the anode may be attributed to the co-activation of lattice oxygen in TiO2-NT and SnO2 by coupled Ru, resulting in a significant amount of •O2 and oxygen vacancies as active sites for glyphosate degradation. After electrolysis, small molecular acids and inorganic ions were obtained, with hydroxylation and dephosphorization as the main degradation pathways. Eight cycles of experiments confirmed that Ru doping prominently enhanced the stability of the GF/PDA/TiO2-NT/SnO2/Ru anode due to its high oxygenophilicity and electron-rich ability, which promoted the generation and utilization efficiency of active free radicals and defects-associated oxygen. Therefore, this study introduces an effective strategy for efficiently co-activating lattice oxygen in SnO2 and TiO2-NT on graphite felt to eliminate persistent organophosphorus pesticides. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for preparation different electrodes.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru, (<b>b</b>) GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>, (<b>c</b>) GF/PDA/TiO<sub>2</sub>-NT/Ru, (<b>d</b>) GF/PDA/TiO<sub>2</sub>-NT. (<b>e</b>) XRD patterns and (<b>f</b>) water contact angle of the four electrodes.</p>
Full article ">Figure 3
<p>(<b>a</b>) A full-scale XPS spectrum of GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru. XPS spectra of (<b>b</b>) Sn 3d, (<b>c</b>) C 1s and Ru 3d, (<b>d</b>) Ti 2p and O 1s of (<b>e</b>) GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru, (<b>f</b>) GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>, (<b>g</b>) GF/PDA/TiO<sub>2</sub>-NT/Ru, (<b>h</b>) GF/PDA/TiO<sub>2</sub>-NT.</p>
Full article ">Figure 4
<p>Effects of (<b>a</b>) pH, (<b>b</b>) initial concentration of glyphosate, (<b>c</b>) current density, (<b>d</b>) Ru loading on the glyphosate degradation efficiency of GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru electrode.</p>
Full article ">Figure 5
<p>(<b>a</b>) Degradation efficiency, (<b>b</b>) TOC removal rate, (<b>c</b>) production rate of PO<sub>4</sub><sup>3−</sup>, (<b>d</b>) energy consumption on GF/PDA/TiO<sub>2</sub>-NT, GF/PDA/TiO<sub>2</sub>-NT/Ru, GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>, GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru electrodes. (<b>e</b>) Recycle experiments of glyphosate degradation, (<b>f</b>) accelerated lifetime test of GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>, GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru electrodes.</p>
Full article ">Figure 6
<p>Electrochemical characterization of the four electrodes: (<b>a</b>) EIS curves, (<b>b</b>) LSV curves, (<b>c</b>) Tafel plots, (<b>d</b>) CV, (<b>e</b>) C<sub>dl</sub> of GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>, GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru electrodes.</p>
Full article ">Figure 7
<p>(<b>a</b>) EPR tests for •OH and (<b>c</b>) •O<sub>2</sub><sup>−</sup> on different electrode; (<b>b</b>) •OH quenching experiments and (<b>d</b>) •O<sub>2</sub><sup>−</sup> quenching experiments on different electrodes.</p>
Full article ">Figure 8
<p>Comparison of XPS spectra of (<b>a</b>) C 1s and Ru 3d, (<b>b</b>) Sn 3d; (<b>c</b>) Ti 2p, (<b>d</b>) O 1s of GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru electrodes before and after electrolysis.</p>
Full article ">Figure 9
<p>Schematic diagram of the glyphosate degradation on GF/PDA/TiO<sub>2</sub>-NT/SnO<sub>2</sub>/Ru anode in electrocatalytic oxidation process.</p>
Full article ">
11 pages, 3198 KiB  
Article
Mo2TiAlC2 as a Saturable Absorber for a Passively Q-Switched Tm:YAlO3 Laser
by Chen Wang, Tianjie Chen, Zhe Meng, Sujian Niu, Zhaoxue Li and Xining Yang
Nanomaterials 2024, 14(22), 1823; https://doi.org/10.3390/nano14221823 - 14 Nov 2024
Viewed by 324
Abstract
Owing to their remarkable characteristics, two-dimensional (2D) layered, MAX phase materials have garnered significant attention in the field of optoelectronics in recent years. Herein, a novel MAX phase ceramic material (Mo2TiAlC2) was prepared into a saturable absorber (SA) by [...] Read more.
Owing to their remarkable characteristics, two-dimensional (2D) layered, MAX phase materials have garnered significant attention in the field of optoelectronics in recent years. Herein, a novel MAX phase ceramic material (Mo2TiAlC2) was prepared into a saturable absorber (SA) by the spin-coating method for passively Q-switching (PQS), and its nonlinear optical absorption properties were characterized with a Tm:YAlO3 (Tm:YAP) nanosecond laser. The structure characteristics and composition analysis revealed that the Mo2TiAlC2 material exhibits a well-defined and stable structure, with a uniform thin film successfully obtained through spin coating. In this study of a PQS laser by employing a Mo2TiAlC2-based SA, an average output power of 292 mW was achieved when the absorbed pump power was approximately 4.59 W, corresponding to a central output wavelength of 1931.2 nm. Meanwhile, a stable pulse with a duration down to 242.9 ns was observed at a repetition frequency of 47.07 kHz, which is the narrowest pulse width recorded among PQS solid-state lasers using MAX phase materials as SAs. Our findings indicate that the Mo2TiAlC2 MAX phase ceramic material is an excellent modulator and has promising potential for ultrafast nonlinear photonic applications. Full article
(This article belongs to the Special Issue Linear and Nonlinear Optical Properties of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Illustration of the preparation process for the Mo<sub>2</sub>TiAlC<sub>2</sub>-based SA.</p>
Full article ">Figure 2
<p>Morphological characterization of the Mo<sub>2</sub>TiAlC<sub>2</sub> powder. (<b>a</b>–<b>c</b>) SEM images of the Mo<sub>2</sub>TiAlC<sub>2</sub> powder at different magnifications; (<b>d</b>,<b>e</b>) EDS images of the Mo<sub>2</sub>TiAlC<sub>2</sub> powder.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) TEM images of the Mo<sub>2</sub>TiAlC<sub>2</sub> powder at different magnifications.</p>
Full article ">Figure 4
<p>Structural and linear absorption characterization of Mo<sub>2</sub>TiAlC<sub>2</sub> material. (<b>a</b>) XRD image of Mo<sub>2</sub>TiAlC<sub>2</sub> powder; (<b>b</b>) Raman spectrum of Mo<sub>2</sub>TiAlC<sub>2</sub> powder; (<b>c</b>) AFM image of Mo<sub>2</sub>TiAlC<sub>2</sub>-based SA; and (<b>d</b>) linear absorption properties of Mo<sub>2</sub>TiAlC<sub>2</sub>-based SA.</p>
Full article ">Figure 5
<p>The experimental setup of the PQS Tm:YAP laser with the Mo<sub>2</sub>TiAlC<sub>2</sub>-based SA.</p>
Full article ">Figure 6
<p>Output performances of PQS Tm:YAP laser. (<b>a</b>) Output power varies with absorbed pump power; (<b>b</b>) pulse width and repetition frequency as a function of absorbed pump power; (<b>c</b>) stable pulse trains of <span class="html-italic">T</span> = 5% (2.5, 100 μs/div, and 1 ms/div); and (<b>d</b>) output spectrum of laser in CW and PQS modes.</p>
Full article ">Figure 7
<p>(<b>a</b>) The beam quality of the PQS Tm:YAP laser with <span class="html-italic">T</span> = 5%; (<b>b</b>) output power stability of the PQS Tm:YAP laser.</p>
Full article ">
12 pages, 1170 KiB  
Article
An Evaluation of Moderate-Refractive-Index Nanoantennas for Enhancing the Photoluminescence Signal of Quantum Dots
by Rafael Ramos Uña, Braulio García Cámara and Ángela I. Barreda
Nanomaterials 2024, 14(22), 1822; https://doi.org/10.3390/nano14221822 - 14 Nov 2024
Viewed by 328
Abstract
The use of nanostructures to enhance the emission of single-photon sources has attracted some attention in the last decade due to the development of quantum technologies. In particular, the use of metallic and high-refractive-index dielectric materials has been proposed. However, the utility of [...] Read more.
The use of nanostructures to enhance the emission of single-photon sources has attracted some attention in the last decade due to the development of quantum technologies. In particular, the use of metallic and high-refractive-index dielectric materials has been proposed. However, the utility of moderate-refractive-index dielectric nanostructures to achieve more efficient single-photon sources remains unexplored. Here, a systematic comparison of various metallic, high-refractive-index and moderate-refractive-index dielectric nanostructures was performed to optimize the excitation and emission of a CdSe/ZnS single quantum dot in the visible spectral region. Several geometries were evaluated in terms of electric field enhancement and Purcell factor, considering the combination of metallic, high-refractive-index and moderate-refractive-index dielectric materials conforming to homogeneous and hybrid nanoparticle dimers. Our results demonstrate that moderate-refractive-index dielectric nanoparticles can enhance the photoluminescence signal of quantum emitters due to their broader electric and magnetic dipolar resonances compared to high-refractive-index dielectric nanoparticles. However, hybrid combinations of metallic and high-refractive-index dielectric nanostructures offer the largest intensity enhancement and Purcell factors at the excitation and emission wavelengths of the quantum emitter, respectively. The results of this work may find applications in the development of single-photon sources. Full article
(This article belongs to the Section Nanophotonics Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A scheme of the illumination of the dimer by a plane wave propagating along the negative direction of the <span class="html-italic">z</span>-axis, i.e., from the air to the substrate, and linearly polarized parallel to the dimer orientation (<span class="html-italic">x</span>-axis). (<b>b</b>) A scheme of the dimer when it is excited by an electric dipole oscillating in the <span class="html-italic">x</span>-direction, representing the QD, located in the middle of the gap and 5 nm above the substrate to take into account the QD size.</p>
Full article ">Figure 2
<p>Scheme of all different dimer configurations. In (<b>a</b>–<b>c</b>), non-hybrid (homogeneous) dimers are represented: (<b>a</b>) two gold NPs, (<b>b</b>) two silicon NPs and (<b>c</b>) two MRI NPs. And in (<b>d</b>–<b>f</b>), hybrid dimers combining all three previous materials are shown. The red dot between the NPs represents the CdSe/ZnS QD.</p>
Full article ">Figure 3
<p>The scattering efficiency and multipolar decomposition for metallic and dielectric isolated NPs in air. The nanoparticles were illuminated by a plane wave propagating along the negative direction of the <span class="html-italic">z</span>-axis (parallel to the disk axis) and linearly polarized along the <span class="html-italic">x</span>-axis (perpendicular to the disk axis). (<b>a</b>) A gold NP of <span class="html-italic">R</span> = 50 nm and <span class="html-italic">H</span> = 150 nm, (<b>b</b>) a silicon NP of <span class="html-italic">R</span> = 120 nm and <span class="html-italic">H</span> = 80 nm and (<b>c</b>) a MRI NP of <span class="html-italic">R</span> = 150 nm and <span class="html-italic">H</span> = 200 nm. Black dotted line: the scattering efficiency obtained with COMSOL. Purple solid line: the scattering efficiency obtained as the sum of the dipolar electric and magnetic, and quadrupolar electric and magnetic, contributions. Blue and red solid lines: dipolar electric and magnetic contributions, respectively. Green and yellow solid lines: quadrupolar electric and magnetic contributions, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) Scattering efficiency for homogeneous dimers: (<b>a</b>) Au-Au, (<b>b</b>) MRI-MRI and (<b>c</b>) Si-Si dimers. The gray vertical lines correspond to the excitation and emission wavelengths. (<b>d</b>–<b>f</b>) Near-field maps (Z-X plane) at the excitation wavelength (570 nm) for the homogeneous dimers: (<b>d</b>) Au-Au, (<b>e</b>) MRI-MRI and (<b>f</b>) Si-Si. A hot-spot is observed in between the NPs, specified as the reddish/greenish area. The dashed white line represents the position of the substrate.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) Scattering efficiency for hybrid dimers: (<b>a</b>) MRI-Au, (<b>b</b>) Si-Au and (<b>c</b>) MRI-Si dimers. The gray vertical lines correspond to the excitation and emission wavelengths. (<b>d</b>–<b>f</b>) Near-field maps (Z-X plane) at the excitation wavelength (570 nm) for the hybrid dimers: (<b>d</b>) MRI-Au, (<b>e</b>) Si-Au and (<b>f</b>) MRI-Si. The dashed white line represents the position of the substrate.</p>
Full article ">
12 pages, 2903 KiB  
Article
Design of Thermo-Responsive Pervaporation Membrane Based on Hyperbranched Polyglycerols and Elastin-like Protein Conjugates
by Juliet Kallon, John J. Bang, Ufana Riaz and Darlene K. Taylor
Nanomaterials 2024, 14(22), 1821; https://doi.org/10.3390/nano14221821 - 14 Nov 2024
Viewed by 338
Abstract
This paper reports the development of a highly crosslinked hyper-branched polyglycerol (HPG) polymer bound to elastin-like proteins (ELPs) to create a membrane that undergoes a distinct closed-to-open permeation transition at 32 °C. The crosslinked HPG forms a robust, mesoporous structure (150–300 nm pores), [...] Read more.
This paper reports the development of a highly crosslinked hyper-branched polyglycerol (HPG) polymer bound to elastin-like proteins (ELPs) to create a membrane that undergoes a distinct closed-to-open permeation transition at 32 °C. The crosslinked HPG forms a robust, mesoporous structure (150–300 nm pores), suitable for selective filtration. The membranes were characterized by FTIR, UV–visible spectroscopy, SEM, and AFM, revealing their structural and morphological properties. Incorporating a synthetic polypeptide introduced thermo-responsive behavior, with the membrane transitioning from impermeable to permeable above the lower critical solution temperature (LCST) of 32 °C. Permeation studies using crystal violet (CV) demonstrated selective transport, where CV permeated only above 32 °C, while water permeated at all temperatures. This hybrid HPG-ELP membrane system, acting as a molecular switch, offers potential for applications in drug delivery, bioseparations, and smart filtration systems, where permeability can be controlled by temperature. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Nuclear magnetic resonance spectra of HPG and methacrylate-functionalized HPG samples. Note the <sup>1</sup>H NMR spectra of HPG displays no peaks in the 5.5–6.5 ppm, denoting the lack of methacrylate protons in this spectrum. These peaks are present in HPG-10, HPG-15, and HPG-28 <sup>1</sup>H NMR spectra.</p>
Full article ">Figure 2
<p>Absorbance of ELP 4-80 as a function of temperature. The UV–Vis spectra show a sharp increase at 33 and 34 °C for 50 µM and 200 µM solutions, respectively, indicating the occurrence of the phase transition.</p>
Full article ">Figure 3
<p>SEM micrographs of HPG-28 lateral cross-sections observed at 40 micron (<b>a</b>), cross-sections observed at 10 micron (<b>b</b>), and edge side view when mixed with 200 µM ELP 4-80 (<b>c</b>). A magnified view of edge of the ELP + HPG-28 membrane reveals less distinct pores compared to HPG-28 alone, giving the appearance that the ELP chain has occupied the main structure of the pores created by HPG-28. (<b>d</b>) Enlargment of cross section shows a uniform and solid, homogenous structure.</p>
Full article ">Figure 4
<p>AFM images on mica of (<b>a</b>) HPG-28 and (<b>b</b>) HPG-15 (experimental repeats, n = 3) membranes. Height distribution of pores for (<b>c</b>) HPG-28 and (<b>d</b>) HPG-15.</p>
Full article ">Figure 5
<p>Permeation of crystal violet solution (400 μL) through ELP 4-80, ELP + HPG-28 chemically crosslinked, ELP + HPG-28 photochemically crosslinked membranes. “No permeation” denotes that no water was observed to permeate through the filter. Each data point was obtained after 3 min of centrifugation at the indicated centrifugal force.</p>
Full article ">Scheme 1
<p>Preparation of functionalized hyperbranched polyglycerols (HPGs) crosslinked to develop pores that incorporate elastin-like proteins.</p>
Full article ">
4 pages, 169 KiB  
Editorial
Nanomaterials in Smart Energy-Efficient Coatings
by Xun Cao
Nanomaterials 2024, 14(22), 1820; https://doi.org/10.3390/nano14221820 - 13 Nov 2024
Viewed by 383
Abstract
Temperature is a key manifestation of energy, with about 51% of global energy consumption occurring in the form of heat annually [...] Full article
(This article belongs to the Special Issue Nanomaterials in Smart Energy-Efficient Coatings)
Previous Issue
Back to TopTop