[go: up one dir, main page]

 
 
applsci-logo

Journal Browser

Journal Browser

Novel Nanomaterials and Nanostructures

A special issue of Applied Sciences (ISSN 2076-3417). This special issue belongs to the section "Materials Science and Engineering".

Deadline for manuscript submissions: closed (30 December 2023) | Viewed by 38212

Special Issue Editors


E-Mail Website
Guest Editor
The Laboratory of Metals and Alloys Under Extreme Impacts, Ufa University of Science and Technology, 450076 Ufa, Russia
Interests: DFT; environmental stability; 2D materials; heterostructures

E-Mail Website
Guest Editor
The Laboratory of Metals and Alloys under Extreme Impacts, Ufa University of Science and Technology, 450076 Ufa, Russia
Interests: atomistic modelling; metals; alloys; nanomaterials

Special Issue Information

Dear Colleagues,

The rise of nanomaterials and nanostructures is fueled by their prediction and discovery using computational and experimental approaches, revealing their wide structural and compositional diversity. The invention of unexplored nanomaterials and nanostructures and the crystal structure modification of pre-existing ones have become the most rapidly developing research directions in modern sciences and industrial projects. This poses new challenges in the development and application of novel nanostructured materials.

This Special Issue is intended for the presentation of new theoretical and experimental results related to the design, chemistry, functional properties and application of novel nanomaterials and nanostructures. Therefore, the overall aim of this Special Issue is to publish high-quality, original research papers in the overlapping fields of:

  • Nanomaterials and nanotechnology
  • Computational materials science
  • Materials synthesis and manufacturing
  • Engineering applications
  • Energy materials
  • Materials characterization
  • Materials degradation
  • Materials chemistry
  • Nanostructured metals and alloys

We look forward to receiving your contributions!

Dr. Andrey Kistanov
Dr. Elena Korznikova
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Applied Sciences is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2400 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • nanostructure design
  • functional properties
  • computer modelling
  • characterization and testing
  • functional materials

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (18 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

12 pages, 4507 KiB  
Article
Transmission and Reflection Spectra of a Bragg Microcavity Filled with a Periodic Graphene-Containing Structure
by Irina V. Fedorova, Svetlana V. Eliseeva and Dmitrij I. Sementsov
Appl. Sci. 2023, 13(13), 7559; https://doi.org/10.3390/app13137559 - 27 Jun 2023
Cited by 1 | Viewed by 1023
Abstract
The transmission and reflection spectra of a one-dimensional microresonator structure with dielectric Bragg mirrors, the working cavity of which is filled with several “dielectric-graphene” or “semiconductor-graphene” periods with controlled material parameters, were obtained using transfer matrices and numerical methods. Carrier drift in graphene [...] Read more.
The transmission and reflection spectra of a one-dimensional microresonator structure with dielectric Bragg mirrors, the working cavity of which is filled with several “dielectric-graphene” or “semiconductor-graphene” periods with controlled material parameters, were obtained using transfer matrices and numerical methods. Carrier drift in graphene monolayers is created to achieve amplification, which makes it possible to use the hydrodynamic approximation to represent graphene conductivity in the terahertz range. The transformation of spectra is achieved both by changing the energy state of the graphene monolayers and by changing the external magnetic field. It is shown that amplification is observed in the region where the real part of the conductivity is negative as the chemical potential (Fermi energy) increases, and the coefficients T and R become substantially greater than unity. The results of the work may be of interest to developers of graphene-based controlled photonic devices. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Symmetric MCR structure formed by two BMs and a resonator cavity designed to be filled with a controllable active structure.</p>
Full article ">Figure 2
<p>The distribution over the structure of the modulus square electric field strength of the wave normalized to the amplitude <math display="inline"><semantics> <msub> <mi>E</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math> (solid line) and the profile of the refractive index (dashed line) (<b>a</b>) and the transmission spectrum of the MCR structure (<b>b</b>) with the thickness of the resonator <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mn>3</mn> </msub> <mo>=</mo> <msub> <mi>λ</mi> <mn>0</mn> </msub> <mo>/</mo> <msqrt> <msub> <mi>ε</mi> <mn>3</mn> </msub> </msqrt> </mrow> </semantics></math> filled with air at the parameters of the layers <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>2.1</mn> </mrow> </semantics></math> (<math display="inline"><semantics> <msub> <mi>SiO</mi> <mn>2</mn> </msub> </semantics></math>), <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>4.16</mn> </mrow> </semantics></math> (<math display="inline"><semantics> <msub> <mi>ZrO</mi> <mn>2</mn> </msub> </semantics></math>), <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (air) and <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Frequency dependence <math display="inline"><semantics> <msub> <mi>ε</mi> <mrow> <mi>s</mi> <mo>⊥</mo> </mrow> </msub> </semantics></math> of the <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>−</mo> <mi>I</mi> <mi>n</mi> <mi>S</mi> <mi>b</mi> </mrow> </semantics></math> semiconductor (<b>a</b>) at <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mn>5</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>15</mn> <mo>,</mo> <mn>20</mn> </mrow> </semantics></math> kOe (red, black, blue, green, purple curves, color online only) and frequency dependence of the real and imaginary parts of the graphene conductivity <math display="inline"><semantics> <mrow> <msub> <mi>σ</mi> <mi>g</mi> </msub> <mo>/</mo> <msub> <mi>σ</mi> <mn>0</mn> </msub> </mrow> </semantics></math> (<b>b</b>) at <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> = 0, 100, 200, 300 meV (red, black, blue, green curves, color online only).</p>
Full article ">Figure 4
<p>Transmission and reflection spectra of a microcavity structure with a graphene-dielectric insert for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>5</mn> <mo>,</mo> <mn>7</mn> </mrow> </semantics></math> (<b>left</b> and <b>right</b>, respectively) at <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>F</mi> </msub> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>200</mn> <mo>,</mo> <mn>300</mn> </mrow> </semantics></math> meV (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 5
<p>Transmission and reflection spectra of a microcavity structure with a graphene-semiconductor insert for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> = 0, 100, 200, 300 meV (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 6
<p>Transmission and reflection spectra of a microcavity structure with a graphene-semiconductor insert for <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>2.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>4.16</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mn>5</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>15</mn> <mo>,</mo> <mn>20</mn> </mrow> </semantics></math> kOe (<b>a</b>–<b>e</b>), <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> = 0, 300 meV (graphs on the <b>left</b> and <b>right</b>, respectively).</p>
Full article ">Figure 7
<p>Dependence of the chemical potential of graphene (Fermi energy) <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> on the external electric field strength <math display="inline"><semantics> <msub> <mi>E</mi> <mn>0</mn> </msub> </semantics></math> at the permittivity of the barrier layer <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mi>b</mi> </msub> <mo>=</mo> <mn>17.8</mn> </mrow> </semantics></math>.</p>
Full article ">
22 pages, 5762 KiB  
Article
Noncovalent Adsorption of Single-Stranded and Double-Stranded DNA on the Surface of Gold Nanoparticles
by Ekaterina A. Gorbunova, Anna V. Epanchintseva, Dmitrii V. Pyshnyi and Inna A. Pyshnaya
Appl. Sci. 2023, 13(12), 7324; https://doi.org/10.3390/app13127324 - 20 Jun 2023
Cited by 2 | Viewed by 1532
Abstract
Understanding the patterns of noncovalent adsorption of double-stranded nucleic acids (dsDNA) on gold nanoparticles (GNPs) was the aim of this study. It was found that the high-affinity motifs in DNA can and do act as an “anchor” for the fixation of the whole [...] Read more.
Understanding the patterns of noncovalent adsorption of double-stranded nucleic acids (dsDNA) on gold nanoparticles (GNPs) was the aim of this study. It was found that the high-affinity motifs in DNA can and do act as an “anchor” for the fixation of the whole molecule on the GNP (up to 98 ± 2 single-stranded (ss)DNA molecules per particle with diameter of 13 ± 2 nm). At the same time, the involvement of an “anchor” in the intramolecular DNA interaction can negatively affect the efficiency of the formation of ss(ds)DNA–GNP structures. It has been shown that the interaction of GNP with DNA duplexes is accompanied by their dissociation and competitive adsorption of ssDNAs on GNP, wherein the crucial factor of DNA adsorption efficiency is the intrinsic affinity of ssDNA to GNP. We propose a detailed scheme for the interaction of dsDNA with GNPs, which should be taken into account in studies of this type. Researchers focused on this field should accept the complicated nature of such objects and take into account the many competing processes, including the processes of adsorption and desorption of DNA on gold as well as the formation of secondary structures by individual DNA strands. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of DNAs used in the work, where F is 5′-CTAACTAACGTCATCATATC-3′, and R is 5′-GATATGATGACGTTAGTTAG-3′. The “main” DNA regions are highlighted in gray or black. The red, blue, and green colors indicate the “anchor” DNA regions.</p>
Full article ">Figure 2
<p>Adsorption capacity of GNPs for ssDNAs. Curves of saturation of GNPs by radiolabeled ssDNAs (<b>A</b>), and a typical plot for KL determination in GraphPad Prism (<b>B</b>).</p>
Full article ">Figure 3
<p>The dependence of the capacity of GNPs for ssDNAs R and F (<b>A</b>) or ssDNAs of the L series (<b>B</b>) or of the T series (<b>C</b>) on the temperature of incubation of the ssDNA with GNPs. Modes of adsorption (on GNPs) of the ssDNA oligonucleotides that do not contain (<b>left</b>) or do contain (<b>right</b>) an “anchor” motif (<b>D</b>).</p>
Full article ">Figure 4
<p>Curves of saturation of GNPs with dsDNA at 25 °C incubation temperature: F–R (<b>A</b>), FL–R (<b>B</b>), FL2–R (<b>C</b>), FL4–R (<b>D</b>), FT–R (<b>E</b>), FT2–R (<b>F</b>), FT4–R (<b>G</b>). * Radioactively labeled ssDNA.</p>
Full article ">Figure 5
<p>The dependence of the capacity of GNPs for a mixture of DNAs on the temperature of incubation of the DNAs with GNPs. (<b>A</b>) DNAs of the L series in a mixture with complementary R, (<b>B</b>) DNAs of the T series in a mixture with complementary R, (<b>C</b>) complementary R mixed with DNA of the L series, (<b>D</b>) complementary R mixed with DNA of the T series, (<b>E</b>) competitive T20 mixed with DNA of the L series, and (<b>F</b>) competitive T20 mixed with DNA of the T series. * denotes radioactively labeled DNA.</p>
Full article ">Figure 6
<p>Mobility and charge analysis (agarose gel electrophoresis) of DNA–GNP associates obtained at 25 °C. A scanned image of an agarose gel (<b>left</b>) and its autoradiograph (<b>right</b>) for FT–GNP associates differing in the capacity of one GNP for ssDNA. The black frames indicate the distribution of the upper (top) and lower (bottom) fractions of FT–GNP associates on a gel lane. The boundary between them is indicated by a blue dotted line. The arrow shows the direction of the electric current.</p>
Full article ">Figure 7
<p>Scans of agarose gels and analysis (in GelAnalyzer 19.1 software, «<a href="http://www.gelanalyzer.com" target="_blank">www.gelanalyzer.com</a>» (accessed on 20 February 2022), created by Istvan Lazar Jr., Ph.D., and Istvan Lazar Sr., Ph.D., CSc, March 2022) of the distribution of FL/GNP associates (<b>A</b>) and of FL–R–GNP associates. T: associate incubation temperature, [DNA]: concentration of ssDNA, Charge: the number of internucleotide phosphate groups. Rows FL*/GNP, R*/GNP, FL*–R/GNP, and FL–R*/GNP: capacity of GNPs for respective associates; % (top) and % (bottom): the proportion of the upper and lower fractions of associates of GNPs with DNA. The arrow indicates the highest electrophoretic mobility of DNA/GNP associates (<b>B</b>).</p>
Full article ">Figure 8
<p>Agarose gel electrophoresis of T20/GNP associates and their mixtures with naked GNPs immediately after the mixing (0 h) and after incubation for 24 h. A gel scan (<b>A</b>) and a gel autoradiograph (<b>B</b>). Naked GNPs and initial T20/GNP associates were applied to the gel as controls for the mobility and distribution of the associates.</p>
Full article ">Figure 9
<p>Schematic representation of the competitive processes of formation of inter- and/or intramolecular structures by oligonucleotides R and FLn (where n = 0, 1, 2, or 4) and their association with GNPs at different molar ratios (the capacity of GNPs is denoted as k, m, z, l, and i) in a single-stranded, double-stranded, or internally structured conformation. The orange background highlights these processes involving R, the gray background indicates the processes involving FLn, the yellow color highlights the formation of the FLn–R duplex, and the green background shows processes involving the FLn–R duplex and GNPs. Index h denotes an internally structured (hairpin) conformation of oligonucleotides R or FLn. Dashed lines show some possible competitive processes.</p>
Full article ">
19 pages, 651 KiB  
Article
Mesoscopic Effects of Interfacial Thermal Conductance during Fast Pre-Melting and Melting of Metal Microparticles
by Alexander Minakov and Christoph Schick
Appl. Sci. 2023, 13(12), 7019; https://doi.org/10.3390/app13127019 - 11 Jun 2023
Viewed by 1223
Abstract
Interfacial thermal conductance (ITC) affects heat transfer in many physical phenomena and is an important parameter for various technologies. The article considers the influence of various mesoscopic effects on the ITC, such as the heat transfer through the gas gap, near-field radiative heat [...] Read more.
Interfacial thermal conductance (ITC) affects heat transfer in many physical phenomena and is an important parameter for various technologies. The article considers the influence of various mesoscopic effects on the ITC, such as the heat transfer through the gas gap, near-field radiative heat transfer, and changes in the wetting behavior during melting. Various contributions to the ITC of the liquid-solid interfaces in the processes of fast pre-melting and melting of metal microparticles are studied. The effective distance between materials in contact is a key parameter for determining ITC. This distance changes significantly during phase transformations of materials. An unusual gradual change in ITC recently observed during pre-melting below the melting point of some metals is discussed. The pre-melting process does not occur on the surface but is a volumetric change in the microstructure of the materials. This change in the microstructure during the pre-melting determines the magnitude of the dispersion forces, the effective distance, and the near-field thermal conductance. The knowledge gained can be useful for understanding and optimizing various technological processes, such as laser additive manufacturing. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>ITC components <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>G</mi> </mrow> <mrow> <mi>n</mi> <mi>f</mi> </mrow> </msub> <mo>(</mo> <mi>d</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>G</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> <mo>(</mo> <mi>d</mi> <mo>)</mo> </mrow> </semantics></math> vs. distance <math display="inline"><semantics> <mrow> <mi>d</mi> </mrow> </semantics></math> for flat surfaces in nitrogen gas at <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> </mrow> </semantics></math> 0.6, <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> </mrow> </semantics></math> 10<sup>5</sup> Pa, and <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> </mrow> </semantics></math> 300 K (<b>a</b>), as well as <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> </mrow> </semantics></math> 900 K (<b>b</b>). The near-field component is estimated as <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>G</mi> </mrow> <mrow> <mi>n</mi> <mi>f</mi> </mrow> </msub> <mo>∈</mo> <mo>(</mo> <msubsup> <mrow> <mi>G</mi> </mrow> <mrow> <mi>n</mi> <mi>f</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> <mo>/</mo> <mn>100</mn> <mo>,</mo> <mtext> </mtext> <msubsup> <mrow> <mi>G</mi> </mrow> <mrow> <mi>n</mi> <mi>f</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> <mo>/</mo> <mn>10</mn> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">
16 pages, 1531 KiB  
Article
Steady-State Crack Growth in Nanostructured Quasi-Brittle Materials Governed by Second Gradient Elastodynamics
by Yury Solyaev
Appl. Sci. 2023, 13(10), 6333; https://doi.org/10.3390/app13106333 - 22 May 2023
Cited by 2 | Viewed by 1126
Abstract
The elastodynamic stress field near a crack tip propagating at a constant speed in isotropic quasi-brittle material was investigated, taking into account the strain gradient and inertia gradient effects. An asymptotic solution for a steady-state Mode-I crack was developed within the simplified strain [...] Read more.
The elastodynamic stress field near a crack tip propagating at a constant speed in isotropic quasi-brittle material was investigated, taking into account the strain gradient and inertia gradient effects. An asymptotic solution for a steady-state Mode-I crack was developed within the simplified strain gradient elasticity by using a representation of the general solution in terms of Lamé potentials in the moving framework. It was shown that the derived solution predicts the nonsingular stress state and smooth opening profile for the growing cracks that can be related to the presence of the fracture process zone in the micro-/nanostructured quasi-brittle materials. Note that similar asymptotic solutions have been derived previously only for Mode-III cracks (under antiplane shear loading). Thus, the aim of this study is to show the possibility of analytical assessments on the elastodynamic crack tip fields for in-plane loading within gradient theories. By using the derived solution, we also performed analysis of the angular distribution of stresses and tractions for the moderate speed of cracks. It was shown that the usage of the maximum principal stress criterion within second gradient elastodynamics allows us to describe a directional stability of Mode-I crack growth and an increase in the dynamic fracture toughness with the crack propagation speed that were observed in the experiments with quasi-brittle materials. Therefore, the possibility of the effective application of regularized solutions of strain gradient elasticity for the refined analysis of dynamic fracture processes in the quasi-brittle materials with phenomenological assessments on the cohesive zone effects is shown. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Illustration for the growing crack problem with global and local (moving) coordinate systems. Opening mode under remotely applied loading is considered. Solution is found in the vicinity of a crack tip, where <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>r</mi> <mo>¯</mo> </mover> <mo>≪</mo> <mn>1</mn> </mrow> </semantics></math>, i.e., <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>≪</mo> <mi>l</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Influence of Poisson’s ratio (<b>a</b>) and Mach number (<b>b</b>) of the deformations of small circles around the crack tip in SGET (left) and classical (right) asymptotic steady-state solutions.</p>
Full article ">Figure 3
<p>Influence of dilatational (<b>a</b>) and rotational (<b>b</b>) amplitudes on the deformed state around the tip of crack.</p>
Full article ">Figure 4
<p>Typical dependence of normalized tractions and double tractions on angular coordinate (<b>a</b>) and in-plane distribution of normalized hoop traction <math display="inline"><semantics> <msub> <mover accent="true"> <mi>t</mi> <mo>^</mo> </mover> <mi>θ</mi> </msub> </semantics></math> ((<b>b</b>), crack is shown by thick black line) in SGET asymptotic solution.</p>
Full article ">Figure 5
<p>Angular distribution of normalized hoop traction (<b>a</b>) and radial traction (<b>b</b>) for different values of Mach number and Poisson’s ratio.</p>
Full article ">Figure 6
<p>Angular distribution of normalized stress components <math display="inline"><semantics> <msub> <mover accent="true"> <mi>σ</mi> <mo>^</mo> </mover> <mrow> <mi>θ</mi> <mi>θ</mi> </mrow> </msub> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <msub> <mover accent="true"> <mi>σ</mi> <mo>^</mo> </mover> <mrow> <mi>r</mi> <mi>θ</mi> </mrow> </msub> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <msub> <mover accent="true"> <mi>σ</mi> <mo>^</mo> </mover> <mrow> <mi>r</mi> <mi>r</mi> </mrow> </msub> </semantics></math> (<b>c</b>) evaluated at different distances from the crack tip and for the values of Mach number <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> (solid lines) and <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> (dashed lines).</p>
Full article ">Figure 7
<p>Angular distribution of normalized maximum principal stress (<b>a</b>) and maximum shear stress (<b>b</b>) and distribution of the normalized maximum principal stress along the crack propagation direction (along <math display="inline"><semantics> <msub> <mi>x</mi> <mn>1</mn> </msub> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) for different values of amplitude ratio <math display="inline"><semantics> <msub> <mi>k</mi> <mi>θ</mi> </msub> </semantics></math> (<b>c</b>). The values of Mach number are <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> (solid lines) and <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> (dashed lines).</p>
Full article ">
16 pages, 4310 KiB  
Article
Magnetic and Optical Properties of Natural Diamonds with Subcritical Radiation Damage Induced by Fast Neutrons
by Nikolai A. Poklonski, Andrey A. Khomich, Ivan A. Svito, Sergey A. Vyrko, Olga N. Poklonskaya, Alexander I. Kovalev, Maria V. Kozlova, Roman A. Khmelnitskii and Alexander V. Khomich
Appl. Sci. 2023, 13(10), 6221; https://doi.org/10.3390/app13106221 - 19 May 2023
Cited by 2 | Viewed by 1691
Abstract
Raman spectroscopy and magnetic properties of the natural single crystalline diamonds irradiated with high fluences of fast reactor neutrons have been investigated. Raman spectra transformations were studied in the range from moderate levels up to radiation damage leading to diamond graphitization. The selection [...] Read more.
Raman spectroscopy and magnetic properties of the natural single crystalline diamonds irradiated with high fluences of fast reactor neutrons have been investigated. Raman spectra transformations were studied in the range from moderate levels up to radiation damage leading to diamond graphitization. The selection of fast neutrons irradiated diamonds for magnetic measurements was carried out according to Raman scattering data on the basis of the intensity criterion and the spectral position of the “1640” band. It was found that in natural diamonds irradiated with neutrons with an extremely high subcritical fluence F = 5 × 1020 cm−2, the transition from a diamagnetic to a ferromagnetic state is observed at the Curie–Weiss temperature of ≈150 K. The energy of the exchange magnetic interaction of uncompensated spins is estimated to be ≈1.7 meV. The differential magnetic susceptibility estimated from the measurements of magnetic moment for temperature 2 K in the limit of B ≈ 0 is χdiff ≈ 1.8 × 10−3 SI units. The nature of magnetism in radiation-disordered single-crystal hydrogen- and metal-free natural diamond grains was discussed. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Raman spectra of natural and CVD diamonds unirradiated and irradiated with fast reactor neutrons with fluences <span class="html-italic">F</span> from 3 × 10<sup>18</sup> to 5 × 10<sup>20</sup> cm<sup>−2</sup> (see also [<a href="#B49-applsci-13-06221" class="html-bibr">49</a>]) or uniformly implanted with multiple-energy keV helium ions with a total dose of 1 × 10<sup>17</sup> and 2.5 × 10<sup>17</sup> cm<sup>−2</sup>. All spectra were recorded with excitation at wavelength λ = 473 nm except for the spectra at <span class="html-italic">F</span> = 3 × 10<sup>18</sup> cm<sup>−2</sup>, which was recorded with excitation at λ = 532 nm.</p>
Full article ">Figure 2
<p>Position (left axis) of the “1640” band and ratio “1640”-to-boson peaks (right axis) in Raman spectra of natural diamonds irradiated with fast neutrons or He-implanted as functions of radiation damage. Dashed lines are guides to eyes. Yellow circles show the position of the “1640” band and ratio “1640”-to-boson peaks corresponding to the critical level of radiation damage of diamond.</p>
Full article ">Figure 3
<p>Transformation of the Raman spectra of fast neutron-irradiated natural diamond crystal with fluence of 5 × 10<sup>20</sup> cm<sup>−2</sup> depending on the annealing temperature (1 h at each temperature in the range of 100–1300 °C = 373–1573 K). Spectra were recorded with excitation at λ = 473 nm.</p>
Full article ">Figure 4
<p>Deconvolution of “1640” and second-order bands of the Raman spectra of fast neutron-irradiated (<span class="html-italic">F</span> = 5 × 10<sup>20</sup> cm<sup>−2</sup>) natural diamond after its annealing at 800 °C (see <a href="#applsci-13-06221-f003" class="html-fig">Figure 3</a>). The contours of the Lorentzian form of the “1640” band of the first and second order are shown in color.</p>
Full article ">Figure 5
<p>Magnetic moment <span class="html-italic">M</span> of natural diamond grains (113 pcs) irradiated with fast reactor neutrons (<span class="html-italic">F</span> = 5 × 10<sup>20</sup> cm<sup>−2</sup>). The measurement absolute temperatures (<span class="html-italic">T</span> from 2 to 300 K) are shown in the figure.</p>
Full article ">Figure 6
<p>Differential magnetic susceptibility χ<sub>diff</sub>(<span class="html-italic">B</span>) = (μ<sub>0</sub>/<span class="html-italic">V</span>)d<span class="html-italic">M</span>/d<span class="html-italic">B</span> (in SI units) of natural diamond grains (113 pcs, <span class="html-italic">V</span> ≈ 7.4 mm<sup>3</sup>) irradiated with fast neutrons (<span class="html-italic">F</span> = 5 × 10<sup>20</sup> cm<sup>−2</sup>). The measurement absolute temperatures (<span class="html-italic">T</span> = 2 and 4 K) are shown in the figure.</p>
Full article ">
15 pages, 56529 KiB  
Article
Synthesis and Printing Features of a Hierarchical Nanocomposite Based on Nickel–Cobalt LDH and Carbonate Hydroxide Hydrate as a Supercapacitor Electrode
by Tatiana L. Simonenko, Nikolay P. Simonenko, Philipp Yu. Gorobtsov, Andrey S. Nikitin, Aytan G. Muradova, Yuri M. Tokunov, Stanislav G. Kalinin, Elizaveta P. Simonenko and Nikolay T. Kuznetsov
Appl. Sci. 2023, 13(10), 5844; https://doi.org/10.3390/app13105844 - 9 May 2023
Cited by 4 | Viewed by 2149
Abstract
The hydrothermal synthesis of a hierarchically organized nanocomposite based on nickel–cobalt carbonate hydroxide hydrate of composition M(CO3)0.5(OH)·0.11H2O (where M is Ni2+ and Co2+) and nickel–cobalt layered double hydroxides (NiCo-LDH) was studied. Using synchronous thermal [...] Read more.
The hydrothermal synthesis of a hierarchically organized nanocomposite based on nickel–cobalt carbonate hydroxide hydrate of composition M(CO3)0.5(OH)·0.11H2O (where M is Ni2+ and Co2+) and nickel–cobalt layered double hydroxides (NiCo-LDH) was studied. Using synchronous thermal analysis (TGA/DSC), it was determined that the material retained thermal stability up to 200 °C. The crystal structure of the powder and the set of functional groups in its composition were determined by X-ray diffraction analysis (XRD) and Fourier transform infrared spectroscopy (FTIR). The resulting hierarchically organized nanopowder was employed as a functional ink component for microplotter printing of an electrode film, which is an array of miniature planar structures with a diameter of about 140 μm, on the surface of a nickel-plated steel substrate. Using scanning electron microscopy (SEM), it was established that the main area of the electrode “pixels” represents a thin film of individual nanorods with periodic inclusions of larger hierarchically organized spherical formations. According to atomic force microscopy (AFM) data, the mean square roughness of the material surface was 28 nm. The electrochemical properties of the printed composite film were examined; in particular, the areal specific capacitance at different current densities was calculated, and the electrochemical kinetics of the material was studied by impedance spectroscopy. It was found that the electrode material under study exhibited relatively low Rs and Rct resistance, which indicates active ion transfer at the electrode/electrolyte interface. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Results of synchronous (TGA/DSC) thermal analysis of the obtained powder.</p>
Full article ">Figure 2
<p>XRD patterns of the obtained powder (left—overview and right—for 2<span class="html-italic">θ</span> range 30°–41° with longer signal accumulation; reflex from the nickel–cobalt layered double hydroxides phase is marked by the “*” marker).</p>
Full article ">Figure 3
<p>FTIR spectrum of the obtained powder (the “*” marker indicates Vaseline oil absorption bands).</p>
Full article ">Figure 4
<p>Microstructure of the obtained nanopowder (according to SEM data).</p>
Full article ">Figure 5
<p>The microplotter printing of functional film (<b>a</b>) and the appearance of the resulting miniature planar nanostructure array on a Ni substrate (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 6
<p>Printed film microstructure and element distribution maps (Ni, Co, and O) on its surface.</p>
Full article ">Figure 7
<p>Microstructure of the printed film (according to SEM data).</p>
Full article ">Figure 8
<p>AFM results of the nickel–cobalt carbonate hydroxide hydrate film: topography (<b>a</b>,<b>b</b>), surface potential distribution map (<b>c</b>), capacity gradient distribution map for “probe tip—sample microregion” capacitor (<b>d</b>), and 3D images of the film at different magnifications (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 9
<p>Electrochemical performance of materials in a three-electrode system with 3 M KOH electrolyte: CV curves at different scan rates (<b>a</b>), GCD curves at different current densities (<b>b</b>) and rate capabilities (inset: cycle performance of the film at 0.300 mA/cm<sup>2</sup> for 2000 cycles) (<b>c</b>), and the EIS spectrum for the material under study (inset: enlargement of high-frequency part and equivalent circuit model) (<b>d</b>).</p>
Full article ">
15 pages, 7065 KiB  
Article
Work Function, Sputtering Yield and Microhardness of an Al-Mg Metal-Matrix Nanostructured Composite Obtained with High-Pressure Torsion
by Rinat Kh. Khisamov, Ruslan U. Shayakhmetov, Yulay M. Yumaguzin, Andrey A. Kistanov, Galiia F. Korznikova, Elena A. Korznikova, Konstantin S. Nazarov, Gulnara R. Khalikova, Rasim R. Timiryaev and Radik R. Mulyukov
Appl. Sci. 2023, 13(8), 5007; https://doi.org/10.3390/app13085007 - 16 Apr 2023
Cited by 4 | Viewed by 1865
Abstract
Severe plastic deformation has proven to be a promising method for the in situ manufacturing of metal-matrix composites with improved properties. Recent investigations have revealed a severe mixing of elements, as well as the formation of non-equilibrium intermetallic phases, which are known to [...] Read more.
Severe plastic deformation has proven to be a promising method for the in situ manufacturing of metal-matrix composites with improved properties. Recent investigations have revealed a severe mixing of elements, as well as the formation of non-equilibrium intermetallic phases, which are known to affect physical and mechanical properties. In this work, a multilayered aluminum–magnesium (Al-Mg) nanostructured composite was fabricated using constrained high-pressure torsion (HPT) in a Bridgeman-anvil-type unit. A microstructure investigation and X-ray diffraction analysis allowed us to identify the presence of intermetallic Al3Mg2 and Al12Mg17 phases in the deformed nanostructured composite. The sputtering yield of the Al3Mg2 and Al12Mg17 phases was found to be 2.2 atom/ion and 1.9 at/ion, respectively, which is lower than that of Mg (2.6 at/ion). According to density functional theory (DFT)-based calculations, this is due to the higher surface-binding energy of the intermetallic phases (3.90–4.02 eV with the Al atom removed and 1.53–1.71 eV with the Mg atom removed) compared with pure Al (3.40–3.84 eV) and Mg (1.56–1.57 eV). In addition, DFT calculations were utilized to calculate the work functions (WFs) of pure Al and Mg and the intermetallic Al3Mg2 and Al12Mg17 phases. The WF of the obtained Al-Mg nanostructured composite was found to be 4 eV, which is between the WF value of Al (4.3 eV) and Mg (3.6 eV). The WF of the Al12Mg17 phase was found to be in a range of 3.63–3.75 eV. These results are in close agreement with the experimentally measured WF of the metal matrix composite (MMC). Therefore, an intermetallic alloy based on Al12Mg17 is proposed as a promising cathode material for various gas-discharge devices, while an intermetallic alloy based on Al3Mg2 is suggested as a promising optical- and acoustic-absorbing material. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of Al (a), Mg (b) and Al-Mg nanostructured composites under HPT conditions (c) and after further annealing at 275 °C (d).</p>
Full article ">Figure 2
<p>SEM images (BSE mode) of the Al-Mg nanostructured composite cross-section at different magnifications and element distribution along the line shown at the top of the panel in the form of a relative intensity plot in relation to the scanning distance of the Mg and Al, marked by green and red, respectively.</p>
Full article ">Figure 3
<p>SEM image (SE mode) of the surface of the Al-Mg nanostructured composite after HPT and annealing following ion irradiation.</p>
Full article ">Figure 4
<p>SEM images (SE mode) of the Al-Mg nanostructured composite surface after ion irradiation. Tilt angle is 45°.</p>
Full article ">Figure 5
<p>SEM image combined with an elemental map of the Al (in red) and Mg (in green). The element composition is shown along lines 1–5 at the top of the panel in the form of a relative intensity plot in relation to the scanning distance of the Al and Mg, marked in red and blue, respectively.</p>
Full article ">Figure 6
<p>AFM image (<b>a</b>) and profile (<b>b</b>) of the Al-Mg composite surface after irradiation.</p>
Full article ">Figure 7
<p>Optical image of the Al-Mg nanostructured composite surface before (<b>a</b>) and after (<b>b</b>) ion irradiation.</p>
Full article ">
14 pages, 2476 KiB  
Article
Rapid Synthesis of Silver Nanowires in the Polyol Process with Conventional and Microwave Heating
by Grzegorz Dzido, Aleksandra Smolska and Muhammad Omer Farooq
Appl. Sci. 2023, 13(8), 4963; https://doi.org/10.3390/app13084963 - 14 Apr 2023
Cited by 4 | Viewed by 3590
Abstract
Silver nanowires (AgNWs) represent an excellent material for many advanced applications due to their thermal and electrical properties. However, synthesising materials with the desired characteristics requires knowledge of the parameters affecting their size and an appropriate fabrication method. This paper presents a study [...] Read more.
Silver nanowires (AgNWs) represent an excellent material for many advanced applications due to their thermal and electrical properties. However, synthesising materials with the desired characteristics requires knowledge of the parameters affecting their size and an appropriate fabrication method. This paper presents a study on the synthesis of silver nanowires using the polyol process by conventional and microwave heating. Various polyols (1,2-ethanediol, 1,3-propanediol, 1,3-butanediol, 1,4-butanediol, 1,5-pentanediol) with different viscosities and dielectric properties were used as reductants. It resulted in nanowires with an average diameter of 119–198 nm. It was found that, in contrast to the viscosity and dielectric constant of the alcohol used, the heating method had a limited effect on the average diameter and length value of the final product. The performed studies indicate an optimal strategy for fabricating one-dimensional silver nanostructures using the polyol method. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Single-mode laboratory setup for microwave-assisted synthesis of AgNWs.</p>
Full article ">Figure 2
<p>XRD patterns of the silver structures prepared in the conventional heating process, reductant (<b>a</b>) 1,3-butanediol or (<b>b</b>) 1,3-propanediol, and a microwave-assisted process, reductant (<b>c</b>) 1,3-butanediol or (<b>d</b>) 1,3-propanediol.</p>
Full article ">Figure 3
<p>Williamson–Hall plot for AgNWs synthesised using different reductants under conventional heating conditions.</p>
Full article ">Figure 4
<p>Williamson–Hall plot for AgNWs synthesised using different reductants in a microwave-assisted process.</p>
Full article ">Figure 5
<p>Silver nanowires synthesised in the microwave-assisted process, 1,3-propanediol reducer: (<b>a</b>) SAED pattern of AgNWs, (<b>b</b>) EDS spectrum for AgNWs.</p>
Full article ">Figure 6
<p>A summary of the properties of AgNWs produced with the application of the different reductants in a batch process under conventional heating conditions: (<b>a</b>) 1,2-ethanediol, (<b>b</b>) 1,3-propanediol, (<b>c</b>) 1,3-butanediol, (<b>d</b>) 1,4-butanediol, (<b>e</b>) 1,5-pentanediol. <span class="html-italic">d</span>, <span class="html-italic">l</span>—average diameter, length, respectively; <span class="html-italic">s</span>—standard deviation, <span class="html-italic">α</span>—conversion; scale bar 10 μm.</p>
Full article ">Figure 7
<p>A summary of the properties of AgNWs produced with the application of the different reductants in a batch, microwave-assisted process: (<b>a</b>) 1,2-ethanediol, (<b>b</b>) 1,3-propanediol, (<b>c</b>) 1,3-butanediol, (<b>d</b>) 1,4-butanediol, (<b>e</b>) 1,5-pentanediol. <span class="html-italic">d</span>, <span class="html-italic">l</span>—average diameter, length, respectively; <span class="html-italic">s</span>—standard deviation; <span class="html-italic">α</span>—conversion; scale bar 10 μm.</p>
Full article ">Figure 8
<p>Growth of the AgNWs due to mutual attachments of one-dimensional objects: (<b>a</b>,<b>b</b>,<b>d</b>) different variants of nanowire-nanowire and nanowire-nanorod structures, (<b>c</b>) V-shape junction.</p>
Full article ">
19 pages, 5446 KiB  
Article
Combined Steam and CO2 Reforming of Methane over Ni-Based CeO2-MgO Catalysts: Impacts of Preparation Mode and Pd Addition
by Lyudmila Okhlopkova, Igor Prosvirin, Mikhail Kerzhentsev and Zinfer Ismagilov
Appl. Sci. 2023, 13(8), 4689; https://doi.org/10.3390/app13084689 - 7 Apr 2023
Cited by 2 | Viewed by 1595
Abstract
The sol–gel template technique makes it possible to synthesize a stable and efficient nickel catalyst based on magnesium-modified cerium oxide Ce0.5Mg0.5O1.5 for the combined steam and CO2 reforming of methane. To stabilize dispersed forms of the active [...] Read more.
The sol–gel template technique makes it possible to synthesize a stable and efficient nickel catalyst based on magnesium-modified cerium oxide Ce0.5Mg0.5O1.5 for the combined steam and CO2 reforming of methane. To stabilize dispersed forms of the active component in the matrix of the support, the catalysts were synthesized by changing the support precursor (cerium acetate and chloride), the active component composition (Ni, NiPd) and the method of introducing nanoparticles. The relationship was established between the physicochemical and catalytic characteristics of the samples. The use of cerium acetate as a support precursor provided smaller pore and crystallite sizes of the support, a stabilization of the dispersed forms of the active component, and excellent catalytic characteristics. The introduction of Pd into the Ni nanoparticles (Pd/Ni = 0.03) increased the resistance of the active component to sintering during the reaction, ensuring stable operation for 25 h of operation. The increased stability was due to a higher concentration of defective oxygen, a higher dispersion of bimetallic NiPd nanoparticles, and the Ni clusters strongly interacting with the NiO-MgO solid solution. An efficient and stable Ni0.194Pd0.006Ce0.4Mg0.4O1.4 catalyst for the conversion of CO2 into important chemicals was developed. With the optimal composition and synthesis conditions of the catalyst, the yield of the target products was more than 75%. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Influence of the calcination temperature on the D<sub>pore</sub> prepared from cerium acetate. 1- Ni/CeMg-C-S-O, 2- Ni/CeMgO-A-S-V, 3- Ni/CeMg-A-S-O, 4- Ni/CeMg-A-P-O.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of the as-prepared catalysts calcined at 500 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-O; 3—Ni/CeMg-A-S-V; 4—Ni-Pd/CeMg-A-S-V; 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 3
<p>X-ray diffraction patterns of the catalysts reduced at 800 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-V; 3—Ni/CeMg-A-S-O; 4—Ni-Pd/CeMg-A-S-V; 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 4
<p>Ce3d XPS spectra of the catalysts reduced at 800 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-O; 3—Ni/CeMg-A-S-V-500; 4—Ni-Pd/CeMg-A-S-V, 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 5
<p>O1s X-ray photoelectron spectra of the catalysts reduced at 800 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-O; 3—Ni/CeMg-A-S-V-500; 4—Ni-Pd/CeMg-A-S-V, 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 6
<p>Hydrogen yield during CSMR at 750 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-V; 3—Ni/CeMg-A-S-O; 4—Ni-Pd/CeMg-A-S-V; 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 7
<p>TGA profiles for the catalysts after the CSMR reaction at 750 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-V; 3—Ni/CeMg-A-S-O; 4—Ni-Pd/CeMg-A-S-V; 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 8
<p>X-ray diffraction patterns of the catalysts after the CSMR reaction at 750 °C. 1—Ni/CeMg-C-S-O; 2—Ni/CeMg-A-S-V; 3—Ni/CeMg-A-S-O; 4—Ni-Pd/CeMg-A-S-V; 5—Ni/CeMg-A-P-O.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) HRTEM and (<b>c</b>) TEM images of the spent Ni-Pd/CeMg-A-S-V catalysts.</p>
Full article ">Figure 10
<p>EDX mapping of the (<b>a</b>,<b>b</b>) spent Ni-Pd/CeMg-A-S-V catalysts, (<b>c</b>,<b>d</b>) C elemental map.</p>
Full article ">Figure 11
<p>EDX mapping of the (<b>a</b>) spent Ni-Pd/CeMg-A-S-V, (<b>b</b>) Ni elemental map, (<b>c</b>) Pd elemental map.</p>
Full article ">
16 pages, 5612 KiB  
Article
Gelation in Alginate-Based Magnetic Suspensions Favored by Poor Interaction among Sodium Alginate and Embedded Particles
by Alexander P. Safronov, Elena V. Rusinova, Tatiana V. Terziyan, Yulia S. Zemova, Nadezhda M. Kurilova, Igor. V. Beketov and Andrey Yu. Zubarev
Appl. Sci. 2023, 13(7), 4619; https://doi.org/10.3390/app13074619 - 6 Apr 2023
Cited by 3 | Viewed by 1649
Abstract
Alginate gels are extensively tested in biomedical applications for tissue regeneration and engineering. In this regard, the modification of alginate gels and solutions with dispersed magnetic particles gives extra options to control the rheo-elastic properties both for the fluidic and gel forms of [...] Read more.
Alginate gels are extensively tested in biomedical applications for tissue regeneration and engineering. In this regard, the modification of alginate gels and solutions with dispersed magnetic particles gives extra options to control the rheo-elastic properties both for the fluidic and gel forms of alginate. Rheological properties of magnetic suspensions based on Na-alginate water solution with embedded magnetic particles were studied with respect to the interfacial adhesion of alginate polymer to the surface of particles. Particles of magnetite (Fe3O4), metallic iron (Fe), metallic nickel (Ni), and metallic nickel with a deposited carbon layer (Ni@C) were taken into consideration. Storage modulus, loss modulus, and the shift angle between the stress and the strain were characterized by the dynamic mechanical analysis in the oscillatory mode. The intensity of molecular interactions between alginate and the surface of the particles was characterized by the enthalpy of adhesion which was determined from calorimetric measurements using a thermodynamic cycle. Strong interaction at the surface of the particles resulted in the dominance of the “fluidic” rheological properties: the prevalence of the loss modulus over the storage modulus and the high value of the shift angle. Meanwhile, poor interaction of alginate polymer with the surface of the embedded particles favored the “elastic” gel-like properties with the dominance of the storage modulus over the loss modulus and low values of the shift angle. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>TEM microphotographs of magnetic particles: (<b>a</b>)—magnetite (Fe<sub>3</sub>O<sub>4</sub>); (<b>b</b>)—metallic iron (Fe); (<b>c</b>)—metallic nickel (Ni); (<b>d</b>)—nickel particles coated with carbon shell (Ni@C); (<b>e</b>)—HRTEM image of the carbon layer on the surface of Ni particle.</p>
Full article ">Figure 2
<p>Shear stress dependence of the dynamic mechanical parameters for Na-alginate water solutions. Frequency of shear oscillations 1Hz, with a temperature of 25 °C. Concentration 5% (wt.): 1—storage modulus, <span class="html-italic">G′</span>; 2—loss modulus, <span class="html-italic">G</span>″; 3—shift angle, <span class="html-italic">δ</span>. Concentration 10% (wt.): 4—storage modulus, <span class="html-italic">G′</span>; 5—loss modulus, <span class="html-italic">G</span>″; 6—shift angle, <span class="html-italic">δ</span>.</p>
Full article ">Figure 3
<p>Concentration dependences of the dynamic moduli and shift angle at 1 Hz frequency and shear stress 1 Pa.</p>
Full article ">Figure 4
<p>Shear stress dependence of the storage modulus (<span class="html-italic">G</span>′), the loss modulus (<span class="html-italic">G</span>″), and the shift angle (<span class="html-italic">δ</span>) for the magneto-rheological suspension of 2.8% Fe<sub>3</sub>O<sub>4</sub> MPs dispersed in 5% Na-alginate solution. Frequency of shear oscillations 1 Hz, temperature 25 °C.</p>
Full article ">Figure 5
<p>Concentration dependences of storage modulus <span class="html-italic">G</span>′ and loss modulus <span class="html-italic">G</span>″ for the magneto-rheological suspensions with different magnetic particles. (<b>a</b>)—Fe and Fe<sub>3</sub>O<sub>4</sub> MPs, (<b>b</b>)—Ni and Ni@C MPs. <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″ are averaged over the plateau in 0.1–1 Pa range of shear stress. Frequency 1 Hz. Temperature 25 °C.</p>
Full article ">Figure 6
<p>Concentration dependences of shift angle <span class="html-italic">δ</span> for magneto-rheological suspensions with different magnetic particles. <span class="html-italic">δ</span> is averaged over the plateau at 0.1–1 Pa range of shear stress. Frequency 1 Hz. Temperature 25 °C.</p>
Full article ">Figure 7
<p>Enthalpy of dissolution of Na-alginate + MPs composites in water (Δ<span class="html-italic">H</span><sub>4</sub>) at 25 °C. Values at the ordinate axes correspond to the enthalpy of dissolution of Na-alginate (left axis—Δ<span class="html-italic">H</span><sub>1</sub>) and to the enthalpy of wetting of MNPs (right axis—Δ<span class="html-italic">H</span><sub>2</sub>). Experimental data for Ni and Ni@C are presented.</p>
Full article ">Figure 8
<p>Concentration dependences of the enthalpy of adhesion of Na-alginate polymer to the surface of MNPs. Temperature 25 °C. Lines are drawn for the eye-guide only.</p>
Full article ">Figure 9
<p>Schematic presentation of the two types of structuring in Na-alginate MRS with MPs depended on the interfacial adhesion among the particles and the polymer. (<b>a</b>)—strong adhesion between Na-alginate and MPs leads to the polymeric layers at the surface which favor fluid-type behavior. It is the case of Fe<sub>3</sub>O<sub>4</sub> and Ni@C MNPs. (<b>b</b>)—weak interaction at the interface leads to the networking of MPs due to the magnetic forces which promote elastic deformation and the gelation of MRS. It is the case of Fe and Ni MPs.</p>
Full article ">
12 pages, 35888 KiB  
Article
Effect of Segregation on Deformation Behaviour of Nanoscale CoCrCuFeNi High-Entropy Alloy
by Arseny M. Kazakov, Azat V. Yakhin, Elvir Z. Karimov, Rita I. Babicheva, Andrey A. Kistanov and Elena A. Korznikova
Appl. Sci. 2023, 13(6), 4013; https://doi.org/10.3390/app13064013 - 21 Mar 2023
Cited by 5 | Viewed by 1842
Abstract
A molecular dynamics (MD) simulation method is used to investigate the effect of grain boundary (GB) segregation on the deformation behavior of bicrystals of equiatomic nanoscale CoCrCuFeNi high-entropy alloy (HEA). The deformation mechanisms during shear and tensile deformation at 300 K and 100 [...] Read more.
A molecular dynamics (MD) simulation method is used to investigate the effect of grain boundary (GB) segregation on the deformation behavior of bicrystals of equiatomic nanoscale CoCrCuFeNi high-entropy alloy (HEA). The deformation mechanisms during shear and tensile deformation at 300 K and 100 K are analyzed. It is revealed that upon tensile deformation, the stacking fault formation, and twinning are the main deformation mechanisms, while for the shear deformation, the main contribution to the plastic flow is realized through the GB migration. The presence of the segregation at GBs leads to the stabilization of GBs, while during the shear deformation of the nanoscale CoCrCuFeNi HEA without the segregation at GBs, GBs are subject to migration. It is found that the GB segregation can differently influence the plasticity of the nanoscale CoCrCuFeNi HEA, depending on the elemental composition of the segregation layer. In the case of copper and nickel segregations, an increase in the segregation layer size enhances the plasticity of the nanoscale CoCrCuFeNi HEA. However, an increase in the thickness of chromium segregations deteriorates the plasticity while enhancing maximum shear stress. The results obtained in this study shed light on the development of HEAs with enhanced mechanical properties via GB engineering. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The computational cell used for the simulations. (<b>b</b>) The bicrystal with the GB segregation of copper (colored in purple) and a crack in the middle of the GB.</p>
Full article ">Figure 2
<p>(<b>a</b>) The shear stress–strain curves for the considered bicrystals. The corresponding type of segregated element is indicated in the legend for each bicrystal, while the case without GB segregation is denoted as Bicrystal. (<b>b</b>) Structure of the bicrystal with GB segregation of Cr at 12% shear deformation.</p>
Full article ">Figure 3
<p>The shear stress–strain curves for the bicrystals with Cr segregation at GB. The corresponding thickness of the segregated layer (1, 3 or 5 Å) is indicated in the legend. The considered temperature is 300 K.</p>
Full article ">Figure 4
<p>The tensile stress–strain curves for the considered bicrystals. The corresponding type of segregated element is indicated in the legend for each bicrystal, while the case without GB segregation is labeled as Bicrystal. The stress drop to zero corresponds to a break in the bicrystal. The considered temperature is 300 K.</p>
Full article ">Figure 5
<p>Bicrystal with chromium segregation at 12% strain. Twins are shown in red. Atoms of the disordered structure are shown in grey.</p>
Full article ">Figure 6
<p>The tensile stress–strain curves for the considered bicrystals. The corresponding type of segregated element is indicated in the legend for each bicrystal, while the case without GB segregation is denoted as Bicrystal. The stress drop to zero corresponds to a break in the bicrystal. The considered temperature is 300 K.</p>
Full article ">Figure 7
<p>Bicrystals during stretching before breaking. Twins are shown in red, and atoms of the disordered structure are shown in grey. The figure shows the bicrystal with cobalt segregation (<b>a</b>), chromium segregation (<b>b</b>), copper segregation (<b>c</b>), iron segregation (<b>d</b>), and nickel segregation (<b>e</b>), and bicrystal without segregations (<b>f</b>).</p>
Full article ">Figure 8
<p>The tensile stress–strain curves for the considered bicrystals. The corresponding type of segregated element is indicated in the legend for each bicrystal, while the case without GB segregation is denoted as Bicrystal. The stress drop to zero corresponds to a break in the bicrystal. The considered temperature is 100 K. The thickness of the segregated layer is 5 Å.</p>
Full article ">
14 pages, 508 KiB  
Article
Characteristics of Resonant Tunneling in Nanostructures with Spacer Layers
by Konstantin Grishakov, Konstantin Katin and Mikhail Maslov
Appl. Sci. 2023, 13(5), 3007; https://doi.org/10.3390/app13053007 - 26 Feb 2023
Cited by 4 | Viewed by 1717
Abstract
The effect of spacer layers on electron transport through two-barrier nanostructures was studied using the numerical solution of the time-dependent Schrodinger–Poisson equations with exact discrete open boundary conditions. The formulation of the problem took into account both the active region consisting of a [...] Read more.
The effect of spacer layers on electron transport through two-barrier nanostructures was studied using the numerical solution of the time-dependent Schrodinger–Poisson equations with exact discrete open boundary conditions. The formulation of the problem took into account both the active region consisting of a quantum well and barriers, as well as the presence of highly doped contact layers and spacer layers. The use of the time formulation of the problem avoids the divergence of the numerical solution, which is usually observed when solving a stationary system of the Schrodinger–Poisson equations at small sizes of spacer layers. It is shown that an increase in the thickness of the emitter spacer leads to a decrease in the peak current through the resonant tunneling nanostructures. This is due to the charge accumulation effects, which, in particular, lead to a change in the potential in an additional quantum well formed in the emitter spacer region when a constant electric field is applied. The valley current also decreases as the thickness of the emitter spacer increases. The peak current and valley current are weakly dependent on the thickness of the collector spacer. The collector spacer thickness has a strong effect on the applied peak and valley voltages. The above features are valid for all three different resonant tunneling nanostructures considered in this study. For the RTD structures based on Al0.3Ga0.7As/GaAs, the optimized peak current value Ipmax = 5.6 × 109 A/m2 and the corresponding applied voltage Vp = 0.44 V. For the RTD structures based on AlAs/In0.8Ga0.2As, Ipmax = 14.5 × 109 A/m2 (Vp = 0.54 V); for RTD structures based on AlAs/In0.53Ga0.47As, Ipmax = 45.5 × 109 A/m2 (Vp = 1.75 V). Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Scheme of the potential profile of the two-barrier resonant tunneling nanostructure and the characteristic dependence of the current on the applied voltage for such a structure. Strongly doped contact layers with donor concentration <math display="inline"><semantics> <msub> <mi>n</mi> <mi>D</mi> </msub> </semantics></math> and thickness <math display="inline"><semantics> <msub> <mi>L</mi> <mi>c</mi> </msub> </semantics></math> are indicated in green; <math display="inline"><semantics> <msub> <mi>I</mi> <mi>p</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>V</mi> <mi>p</mi> </msub> </semantics></math> are the peak current and peak voltage; <math display="inline"><semantics> <msub> <mi>I</mi> <mi>v</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>V</mi> <mi>v</mi> </msub> </semantics></math> are the valley current and valley voltage. <math display="inline"><semantics> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> </semantics></math>—thickness of emitter spacer layer; <math display="inline"><semantics> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>C</mi> </mrow> </msub> </semantics></math>—thickness of collector spacer layer; <math display="inline"><semantics> <msub> <mi>L</mi> <mi>b</mi> </msub> </semantics></math>—thickness of barrier layer; <math display="inline"><semantics> <msub> <mi>L</mi> <mrow> <mi>Q</mi> <mi>W</mi> </mrow> </msub> </semantics></math>—thickness of quantum well; <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>E</mi> <mi>c</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msubsup> <mi>E</mi> <mi>c</mi> <mrow> <mi>Q</mi> <mi>W</mi> </mrow> </msubsup> </mrow> </semantics></math>—conduction band discontinuities between layers with different band gaps.</p>
Full article ">Figure 2
<p>The scheme of calculation of the static current–voltage characteristic based on the solution of the time-dependent Schrodinger–Poisson equations. <b>Left</b> graph: the dependence of the applied voltage on time; <b>central</b> graph: the dependence of the current through the nanostructure on time; <b>right</b> graph: the dependence of the direct current on voltage.</p>
Full article ">Figure 3
<p>Current–voltage characteristic for the resonant tunneling nanostructure from [<a href="#B23-applsci-13-03007" class="html-bibr">23</a>]. The solid blue line is the solution from [<a href="#B23-applsci-13-03007" class="html-bibr">23</a>], and the red dots represent the solution obtained as a result of our calculations.</p>
Full article ">Figure 4
<p>Time dependence of the value <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>ρ</mi> <mn>0</mn> </msub> <mo>/</mo> <msub> <mi>ρ</mi> <mi>D</mi> </msub> <mo>=</mo> <mfrac> <mrow> <mo>∫</mo> <mfenced separators="" open="(" close=")"> <mi>n</mi> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>−</mo> <msub> <mi>n</mi> <mi>D</mi> </msub> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </mfenced> <mi>d</mi> <mi>x</mi> </mrow> <mrow> <mo>∫</mo> <msub> <mi>n</mi> <mi>D</mi> </msub> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mi>d</mi> <mi>x</mi> </mrow> </mfrac> </mrow> </semantics></math> characterizing the fulfillment of the electroneutrality condition.</p>
Full article ">Figure 5
<p>The dependence of current (left <span class="html-italic">y</span>-axis) and voltage (right <span class="html-italic">y</span>-axis) on time for Nanostructure #1 at <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> nm.</p>
Full article ">Figure 6
<p>The dependence of the peak current on the thickness of the emitter spacer calculated in the simplified model (without the accumulation charge effect and highly doped contact layers) for the considered resonant tunneling nanostructures.</p>
Full article ">Figure 7
<p>The dependence of the peak current on the thickness of the emitter spacer for the considered resonant tunneling nanostructures.</p>
Full article ">Figure 8
<p>(<b>Left</b>) The potential of the bottom of the conduction band (right <span class="html-italic">y</span>-axis, solid curve) and electron density (left <span class="html-italic">y</span>-axis, dotted curve) depending on the coordinate for the resonant tunneling Nanostructure #1 at the applied voltage <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mrow> <mi>d</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>0.29</mn> </mrow> </semantics></math> V. The position of the Fermi level is shown by a green fill. (<b>Right</b>) The potential of the bottom of the conduction band for Nanostructure #1, maximized in the emitter spacer region, for several values of the spacer thickness.</p>
Full article ">Figure 9
<p>Current–voltage characteristics of the RTD Structure #2 at various emitter spacer thicknesses: <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> nm—black lines, <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> nm—blue lines, <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> nm—red lines, <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>E</mi> </mrow> </msub> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math> nm—green lines. Solid lines correspond to <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>C</mi> </mrow> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> nm, and dashed lines correspond to <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>C</mi> </mrow> </msub> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math> nm.</p>
Full article ">Figure 10
<p>Dependences of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>I</mi> <mo>=</mo> <msub> <mi>I</mi> <mi>p</mi> </msub> <mo>−</mo> <msub> <mi>I</mi> <mi>v</mi> </msub> </mrow> </semantics></math> (red line, left <span class="html-italic">Y</span>-axis) and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>V</mi> <mo>=</mo> <msub> <mi>V</mi> <mi>v</mi> </msub> <mo>−</mo> <msub> <mi>V</mi> <mi>p</mi> </msub> </mrow> </semantics></math> (blue line, right <span class="html-italic">Y</span>-axis) on the thickness of the emitter spacer for RTD Structure #1 at <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mrow> <mi>S</mi> <mi>C</mi> </mrow> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> nm.</p>
Full article ">
9 pages, 3057 KiB  
Article
Nonlinear Transport and Magnetic/Magneto-Optical Properties of Cox(MgF2)100-x Nanostructures
by Sergey A. Ivkov, Konstantin A. Barkov, Evelina P. Domashevskaya, Elena A. Ganshina, Dmitry L. Goloshchapov, Stanislav V. Ryabtsev, Alexander V. Sitnikov and Pavel V. Seredin
Appl. Sci. 2023, 13(5), 2992; https://doi.org/10.3390/app13052992 - 26 Feb 2023
Cited by 3 | Viewed by 1238
Abstract
The aim of this work was to comprehensively study the effect of the variable atomic composition and structural-phase state of Cox(MgF2)100-x nanocomposites on their nonlinear transport and magnetic/magneto-optical properties. Micrometer-thick nanocomposite layers on glass substrates were obtained by [...] Read more.
The aim of this work was to comprehensively study the effect of the variable atomic composition and structural-phase state of Cox(MgF2)100-x nanocomposites on their nonlinear transport and magnetic/magneto-optical properties. Micrometer-thick nanocomposite layers on glass substrates were obtained by means of ion-beam sputtering of a composite target in the argon atmosphere in a wide range of compositions (x = 16–59 at.%). Using a low metal content in the nanocomposite, magnesium fluoride was kept in the nanocrystalline state. As the metal content increased, nanocrystalline cobalt was formed. The value of the resistive percolation threshold, xper = 37 at.%, determined from the concentration dependences of the electrical resistance of the nanocomposites coincided with the beginning of nucleation of the metallic nanocrystals in the MgF2 dielectric matrix. The absolute value of the maximum negative magnetoresistive effect in the nanocomposites was 5% in a magnetic field of 5.5 kG at a Co concentration of x = 27 at.%. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of Co<sub>x</sub>(MgF<sub>2</sub>)<sub>100–x</sub> nanocomposites of various compositions on glass substrates [<a href="#B25-applsci-13-02992" class="html-bibr">25</a>].</p>
Full article ">Figure 2
<p>Concentration dependences of electrical resistance (<b>a</b>) and magnetoresistance (<b>b</b>) of Co<sub>x</sub>(MgF<sub>2</sub>)<sub>100-x</sub> nanocomposites.</p>
Full article ">Figure 3
<p>Impedance diagrams (hodographs) for nanocomposites Co<sub>x</sub>(MgF<sub>2</sub>)<sub>100-x</sub> of different compositions (32 ≤ x ≤ 59).</p>
Full article ">Figure 4
<p>Concentration dependencies of the transversal Kerr effect (TKE) for nanocomposites Co<sub>x</sub>(MgF<sub>2</sub>)<sub>100-x</sub> at different values of incident light energy.</p>
Full article ">Figure 5
<p>The M–H loop for NC Co<sub>32</sub>(MgF<sub>2</sub>)<sub>68</sub> and for NC Co<sub>27</sub>(MgF<sub>2</sub>)<sub>73</sub>, where the solid delta shows the M–H loop with an out-of-plane magnetic field, and the open delta presents the in-plane M–H loop.</p>
Full article ">Figure 6
<p>Concentration dependence of the coercive force (H<sub>c</sub>) for Co<sub>x</sub>(MgF<sub>2</sub>)<sub>100-x</sub> nanocomposites.</p>
Full article ">
7 pages, 1637 KiB  
Communication
First Theoretical Realization of a Stable Two-Dimensional Boron Fullerene Network
by Bohayra Mortazavi
Appl. Sci. 2023, 13(3), 1672; https://doi.org/10.3390/app13031672 - 28 Jan 2023
Cited by 2 | Viewed by 1982
Abstract
Successful experimental realizations of two-dimensional (2D) C60 fullerene networks have been among the most exciting latest advances in the rapidly growing field of 2D materials. In this short communication, on the basis of the experimentally synthesized full boron B40 fullerene lattice, [...] Read more.
Successful experimental realizations of two-dimensional (2D) C60 fullerene networks have been among the most exciting latest advances in the rapidly growing field of 2D materials. In this short communication, on the basis of the experimentally synthesized full boron B40 fullerene lattice, and by structural minimizations of extensive atomic configurations via density functional theory calculations, we could, for the first time, predict a stable B40 fullerene 2D network, which shows an isotropic structure. Acquired results confirm that the herein predicted B40 fullerene network is energetically and dynamically stable and also exhibits an appealing thermal stability. The elastic modulus and tensile strength are estimated to be 125 and 7.8 N/m, respectively, revealing strong bonding interactions in the predicted nanoporous nanosheet. Electronic structure calculations reveal metallic character and the possibility of a narrow and direct band gap opening by applying the uniaxial loading. This study introduces the first boron fullerene 2D nanoporous network with an isotropic lattice, remarkable stability, and a bright prospect for the experimental realization. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) 3D view of the B<sub>40</sub> cage. (<b>b</b>) Top and side views for the 2D boron fullerene network. (<b>c</b>) 3D view for the electronic localization function (with yellow color) of the 2D boron fullerene network with an isosurface value of 0.75, illustrated using the VESTA package [<a href="#B31-applsci-13-01672" class="html-bibr">31</a>]. Find the energy minimized structures in the <a href="#app1-applsci-13-01672" class="html-app">Supporting Information document</a>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Phonon dispersion and (<b>b</b>) group velocity of the predicted single-layer B<sub>40</sub> network. (<b>c</b>) Uniaxial stress–strain at the ground state along with side views of the deformed structures. (<b>d</b>) The AIMD results for the per atom total energy of the B<sub>40</sub> nanosheet during the simulations at temperatures of 500, 700, and 1000 K. The insets in panel (<b>d</b>) show the side views for the final atomic configurations after 20 ps of AIMD simulations.</p>
Full article ">Figure 3
<p>Electronic band structures of the stress-free and strained B<sub>40</sub> 2D network by the PBE (solid lines) and HSE06 (dotted lines) methods.</p>
Full article ">
13 pages, 33471 KiB  
Article
Effect of Interatomic Potential on Simulation of Fracture Behavior of Cu/Graphene Composite: A Molecular Dynamics Study
by Liliya R. Safina, Elizaveta A. Rozhnova, Ramil T. Murzaev and Julia A. Baimova
Appl. Sci. 2023, 13(2), 916; https://doi.org/10.3390/app13020916 - 9 Jan 2023
Cited by 11 | Viewed by 2266
Abstract
Interatomic interaction potentials are compared using a molecular dynamics modeling method to choose the simplest, but most effective, model to describe the interaction of copper nanoparticles and graphene flakes. Three potentials are considered: (1) the bond-order potential; (2) a hybrid embedded-atom-method and Morse [...] Read more.
Interatomic interaction potentials are compared using a molecular dynamics modeling method to choose the simplest, but most effective, model to describe the interaction of copper nanoparticles and graphene flakes. Three potentials are considered: (1) the bond-order potential; (2) a hybrid embedded-atom-method and Morse potential; and (3) the Morse potential. The interaction is investigated for crumpled graphene filled with copper nanoparticles to determine the possibility of obtaining a composite and the mechanical properties of this material. It is observed that not all potentials can be applied to describe the graphene–copper interaction in such a system. The bond-order potential potential takes into account various characteristics of the bond (for example, the angle of rotation and bond lengths); its application increases the simulation time and results in a strong interconnection between a metal nanoparticle and a graphene flake. The hybrid embedded-atom-method/Morse potential and the Morse potential show different results and lower bonding between graphene and copper. All the potentials enable a composite structure to be obtained; however, the resulting mechanical properties, such as strength, are different. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Graphene flake with Cu nanoparticle inside and (<b>b</b>) composite precursor at the initial state and after exposure at 300 K. Cu atoms are shown in orange and C atoms are shown in black color.</p>
Full article ">Figure 2
<p>(<b>a</b>) Snapshots of the composite precursors obtained with different potentials after exposure at 300 K followed by hydrostatic compression. (<b>b</b>,<b>c</b>) Pressure and potential energy changes during hydrostatic compression. (<b>d</b>) Snapshots of the composite, obtained with different potentials. Cu atoms are shown in orange and C atoms are shown in black.</p>
Full article ">Figure 3
<p>Stress-strain curves for composites under tension. The insertion shows the elastic regime and elastic modulus.</p>
Full article ">Figure 4
<p>Copper atoms (orange color) and graphene network (black color) are shown separately before and after uniaxial tension: (<b>a</b>) BOP; (<b>b</b>) EAM-Morse; and (<b>c</b>) Morse. Colors as in <a href="#applsci-13-00916-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
8 pages, 1206 KiB  
Communication
First-Principles Prediction of Structure and Properties of the Cu2TeO6 Monolayer
by Elena A. Korznikova, Vladimir A. Bryzgalov and Andrey A. Kistanov
Appl. Sci. 2023, 13(2), 815; https://doi.org/10.3390/app13020815 - 6 Jan 2023
Cited by 1 | Viewed by 1857
Abstract
In this work, first-principles calculations have been utilized to predict the existence of a new Cu2TeO6 monolayer. It is shown that the predicted material is dynamically and thermally stable. The Cu2TeO6 monolayer is also found to be [...] Read more.
In this work, first-principles calculations have been utilized to predict the existence of a new Cu2TeO6 monolayer. It is shown that the predicted material is dynamically and thermally stable. The Cu2TeO6 monolayer is also found to be a narrow band gap semiconductor with a band gap size of 0.20 eV. Considering the obtained properties of the Cu2TeO6 monolayer, it is proposed for applications in various nanodevices in electronics and straintronics. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>The atomic structure of the unit cell (<b>a</b>) and phonon dispersion curve (<b>b</b>) for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer.</p>
Full article ">Figure 2
<p>Total energy fluctuation obtained from AIMD simulations conducted at 300 K (<b>a</b>) and 320 K (<b>c</b>) for the time of 4 ps for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer. The atomic structure of the Cu<sub>2</sub>TeO<sub>6</sub> monolayer at 300 K (<b>b</b>) and 320 K (<b>d</b>) after 4 ps.</p>
Full article ">Figure 3
<p>ELF with the isosurface value of 0.70 for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer (<b>a</b>) and WF for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer (<b>b</b>).</p>
Full article ">Figure 4
<p>Band structure (<b>a</b>) and LDOS (<b>b</b>) for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer. The Fermi level is represented by the dashed green line.</p>
Full article ">Figure 5
<p>Spatial dependencies of (<b>a</b>) the Young’s modulus, (<b>b</b>) the shear modulus, and (<b>c</b>) Poisson’s ratio for the Cu<sub>2</sub>TeO<sub>6</sub> monolayer.</p>
Full article ">

Review

Jump to: Research

23 pages, 2912 KiB  
Review
Prospects for Combined Applications of Nanostructured Catalysts and Biocatalysts for Elimination of Hydrocarbon Pollutants
by Olga Maslova, Olga Senko, Marina A. Gladchenko, Sergey N. Gaydamaka and Elena Efremenko
Appl. Sci. 2023, 13(9), 5815; https://doi.org/10.3390/app13095815 - 8 May 2023
Cited by 4 | Viewed by 2049
Abstract
Due to the presence of environmental problems, it is urgent to improve the processes aimed at the processing and purification of hydrocarbon-containing wastes and wastewaters. The review presents the latest achievements in the development of nanostructured catalysts made from different materials that can [...] Read more.
Due to the presence of environmental problems, it is urgent to improve the processes aimed at the processing and purification of hydrocarbon-containing wastes and wastewaters. The review presents the latest achievements in the development of nanostructured catalysts made from different materials that can be used to purify oil-polluted wastewaters (petroleum refinery wastewater, oilfield-produced water, sulfur-containing extracts from pre-oxidized crude oil and oil fractions, etc.) and eliminate components of hydrocarbon pollutants (polyaromatic hydrocarbons, phenols, etc.). The results of the analysis of possible combinations of chemical and biological catalysts for deeper and more effective solutions to the problems are discussed. The possibilities of highly efficient elimination of hydrocarbon pollutants as a result of the hybrid application of nanoparticles (graphene oxide, mesoporous silica, magnetic nanocatalysts, etc.) or catalytic nanocomposites for advanced oxidation processes and biocatalysts (enzymes, cells of bacteria, mycelial fungi, phototrophic microorganisms and natural or artificial microbial consortia) are analyzed. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Different catalysts and processes based on their use in the treatments of oil-containing wastewaters.</p>
Full article ">Figure 2
<p>Various BCs used for the degradation of hydrocarbon pollutants and wastewaters polluted by oil residues discussed in the review. The direct effect on hydrocarbon pollutants is provided by enzymes that are NCs (marked by violet). These enzymes can be synthesized by microbial catalysts directly in the process of destruction, as well as isolated and used as parts of NComs. The outer contour contains a set of the most commonly used microbial BCs: bacteria and consortia (marked by orange), mycelial fungi (marked by blue) and phototrophic microorganisms and those microbial cells that are not effective destructors of hydrocarbon pollutants but act as auxiliary “personnel” in such processes (marked by green).</p>
Full article ">
34 pages, 9921 KiB  
Review
Recent Trends in the Characterization and Application Progress of Nano-Modified Coatings in Corrosion Mitigation of Metals and Alloys
by Abhinay Thakur, Savaş Kaya and Ashish Kumar
Appl. Sci. 2023, 13(2), 730; https://doi.org/10.3390/app13020730 - 4 Jan 2023
Cited by 62 | Viewed by 4116
Abstract
Nanotechnology is a discipline of science and engineering that emphasizes developing, modifying, characterizing, and using nanoscale components in a variety of applications. Owing to their multiple advantages, including adhesion strength, surface hardness, long-term and extra-high-temperature corrosion resistance, improvement of interfacial behavior, etc., nanocoatings [...] Read more.
Nanotechnology is a discipline of science and engineering that emphasizes developing, modifying, characterizing, and using nanoscale components in a variety of applications. Owing to their multiple advantages, including adhesion strength, surface hardness, long-term and extra-high-temperature corrosion resistance, improvement of interfacial behavior, etc., nanocoatings are efficiently utilized to minimize the influence of a corrosive environment. Additionally, nanocoatings are often applied in thinner and finer concentrations, allowing for greater versatility in instrumentation and reduced operating and maintenance costs. The exemplary physical coverage of the coated substrate is facilitated by the fine dimensions of nanomaterials and the significant density of their grounded boundaries. For instance, fabricated self-healing eco-sustainable corrosion inhibitors including PAC/CuONPs, PAC/Fe3O4NPs, and PAC/NiONPs, with uniform distributions and particulate sizes of 23, 10, and 43 nm, correspondingly, were effective in producing PAC/MONPs nanocomposites which exhibited IE% of 93.2, 88.1, 96.1, and 98.6% for carbon steel corrosion in 1M HCl at the optimum concentration of 250 ppm. Therefore, in this review, further steps are taken into the exploration of the significant corrosion-mitigation potential and applications of nanomaterial-based corrosion inhibitors and nano-modified coatings, including self-healing nanocoatings, natural source-based nanocoatings, metal/metallic ion-based nanocoatings, and carbon allotrope-based nanocoatings, to generate defensive film and protection against corrosion for several metals and alloys. These have been illuminated through the in-depth discussion on characterization techniques such as scanning electron microscopy (SEM), electrochemical impedance spectroscopy (EIS), potentiodynamic polarization (PDP), atomic force microscopy (AFM), energy dispersive spectroscopy (EDS), etc. After providing a general summary of the various types of nanomaterials and their protective mechanisms in wide corrosive media, we subsequently present a viewpoint on challenges and future directions. Full article
(This article belongs to the Special Issue Novel Nanomaterials and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>The pictorial representation of infusing Y<sub>2</sub>O<sub>3</sub> using imidazole in the epoxy coating. Reprinted with permission from Ref. [<a href="#B35-applsci-13-00730" class="html-bibr">35</a>]. MDPI (2021).</p>
Full article ">Figure 2
<p>Morphological scans of the surface (3 mm × 3 cm) of the Mg-Al sample; undipped and dipped in an HBSS following 14 days for specimens. Reprinted with permission from Ref. [<a href="#B36-applsci-13-00730" class="html-bibr">36</a>]. MDPI (2022).</p>
Full article ">Figure 3
<p>SEM scans of the substrates: (<b>a</b>) specimen 0, untreated AZ61 alloy; (<b>b</b>) specimen A, deteriorated AZ61 alloy following the HBSS investigation; (<b>c</b>) deteriorated specimen A-SG; (<b>d</b>) deteriorated specimen A-SG-CIS; (<b>e</b>) deteriorated specimen A-SG-DMG; (<b>f</b>) deteriorated specimen A-SG-QUI; (<b>g</b>) deteriorated specimen A-SG-AMIN; (<b>h</b>) deteriorated specimen A-SG-GRA; and (<b>i</b>) deteriorated specimen A-SG-TI. Reprinted with permission from Ref. [<a href="#B36-applsci-13-00730" class="html-bibr">36</a>]. MDPI (2022).</p>
Full article ">Figure 4
<p>SEM scans of CMC.BTA microspheres attained by the technique of spray-drying in distinctive empirical situations: (<b>a</b>) 440 L/h and 170 °C; (<b>b</b>) 600 L/h and 170 °C; (<b>c</b>) 440 L/h and 180 °C; (<b>d</b>) 600 L/h and 180 °C; (<b>e</b>) 440 L/h and 190 °C; and (<b>f</b>) 600 L/h and 190 °C. Reprinted with permission from Ref. [<a href="#B18-applsci-13-00730" class="html-bibr">18</a>]. MDPI (2022).</p>
Full article ">Figure 5
<p>FT-IR spectra of magnetic hydroxyapatite NPs. Reprinted with permission from Ref. [<a href="#B44-applsci-13-00730" class="html-bibr">44</a>]. MDPI (2022).</p>
Full article ">Figure 6
<p>Micrography of four distinct covered substrate and interfacial coatings of steel: (<b>a</b>) pristine Sn covering (S1); (<b>b</b>) Sn + ZnO NPs covering (S2); (<b>c</b>) Sn + NiO NPs covering (S3); and (<b>d</b>) Sn + ZnO + NiO NPs covering (S4). Reprinted with permission from Ref. [<a href="#B45-applsci-13-00730" class="html-bibr">45</a>] MDPI (2022).</p>
Full article ">Figure 7
<p>Nyquist graph for LCS, LCS + Sn, LCS + Sn + ZnO, LCS + Sn + NiO, and LCS + Sn + ZnO + NiO specimens following 1 h dipping in 3.5% NaCl medium. Reprinted with permission from Ref. [<a href="#B45-applsci-13-00730" class="html-bibr">45</a>] MDPI (2022).</p>
Full article ">Figure 8
<p>Impedance curves for C-steel in acidic media: (<b>A</b>) Nyquist with the exclusion and inclusion of various dosages of PAC; (<b>B</b>) Nyquist with the exclusion and inclusion of 250 ppm of several nanocomposite inhibitors; and (<b>C</b>) Bode and (<b>D</b>) phase curve illustration with the exclusion and inclusion of various dosage of PAC at 50 °C. Reprinted with permission from Ref. [<a href="#B74-applsci-13-00730" class="html-bibr">74</a>]. MDPI (2021).</p>
Full article ">Figure 9
<p>Illustration for the green production of Cr<sub>2</sub>O<sub>3</sub> NPs utilizing <span class="html-italic">Abutilon indicum</span> (L.) Sweet leaf extract. Reprinted with permission from Ref. [<a href="#B5-applsci-13-00730" class="html-bibr">5</a>]. MDPI (2021).</p>
Full article ">Figure 10
<p>The optimized molecular configurations, HOMO, and LUMO of both inhibitors using DMol<sup>3</sup> array. Reprinted with permission from Ref. [<a href="#B8-applsci-13-00730" class="html-bibr">8</a>]. MDPI (2021).</p>
Full article ">Figure 11
<p>The optimal arrangement for adsorption of the CEL and NCC over the Fe (1 1 0) surface attained through the adsorption locator array. Reprinted with permission from Ref. [<a href="#B8-applsci-13-00730" class="html-bibr">8</a>]. MDPI (2021).</p>
Full article ">Figure 12
<p>Expedited carbonation for concrete samples: (<b>a</b>) regulated and (<b>b</b>) GA-NP inhibitor. Reprinted with permission from Ref. [<a href="#B83-applsci-13-00730" class="html-bibr">83</a>]. MDPI 2021.</p>
Full article ">Figure 13
<p>AFM scans for steel reinforcement substrates subjected to CO<sub>2</sub> for a duration of 180 days: (<b>a</b>,<b>b</b>) 2D, 3D scans for regulated and (<b>c</b>,<b>d</b>) 2D, 3D scans for inhibited samples including GA-NPs. Reprinted with permission from Ref. [<a href="#B83-applsci-13-00730" class="html-bibr">83</a>]. MDPI 2021.</p>
Full article ">Figure 14
<p>Surface topologies of (<b>a</b>) Ni-P, (<b>b</b>) Ni-P-SN, (<b>c</b>) Ni-P-nSN composite, (<b>d</b>,<b>e</b>) Ni-P-SN and Ni-P-nSN composed of a coating (~5 µm). Reprinted with permission from Ref. [<a href="#B85-applsci-13-00730" class="html-bibr">85</a>]. MDPI (2022).</p>
Full article ">Figure 15
<p>EDS findings of the deteriorated substrate (<b>region I</b>) and original substrate (<b>region II</b>) of (<b>a</b>,<b>b</b>) Ni-P, (<b>c</b>,<b>d</b>) Ni-P-SN, and (<b>e</b>,<b>f</b>) Ni-P-nSN samples. Reprinted with permission from Ref. [<a href="#B85-applsci-13-00730" class="html-bibr">85</a>]. MDPI (2022).</p>
Full article ">Figure 16
<p>AFM scans illustrating significant surface coverage shown by trans 5,15-(4-carboxy-phenyl)-10,20- diphenylporphyrin (<b>a</b>) in contrast to cis configuration, 5,10-(4-carboxy-phenyl)-15,20-(4-phenoxy- phenyl)-porphyrin (<b>b</b>). Reprinted with permission from Ref. [<a href="#B86-applsci-13-00730" class="html-bibr">86</a>]. MDPI 2021.</p>
Full article ">Figure 17
<p>EIS findings of Q235 CS in 1 M HCl medium in the exclusion and inclusion of various dosages of NCDs: (<b>a</b>) Nyquist graph and (<b>b</b>) Bode graph. Reprinted with permission from Ref. [<a href="#B90-applsci-13-00730" class="html-bibr">90</a>]. MDPI (2021).</p>
Full article ">Figure 18
<p>LSCM and SEM scans of Q235 CS substrate following 60 h of subjection to 1 M HCl medium with the exclusion and inclusion of 400 ppm NCDs, (<b>a</b>,<b>b</b>) refined steel, (<b>c</b>,<b>d</b>) in blank media, (<b>e</b>,<b>f</b>) with the inclusion of 400 ppm NCDs. Reprinted with permission from Ref. [<a href="#B90-applsci-13-00730" class="html-bibr">90</a>]. MDPI (2021).</p>
Full article ">Figure 19
<p>Illustration displaying the inhibition mechanism of Y<sub>2</sub>O<sub>3</sub>/IMD-modified covering over the surface of the steel. Reprinted with permission from Ref. [<a href="#B35-applsci-13-00730" class="html-bibr">35</a>]. MDPI (2021).</p>
Full article ">
Back to TopTop