[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (109)

Search Parameters:
Keywords = diamond nanoparticles

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 18048 KiB  
Article
Detonation Nanodiamond Soot—A Structurally Tailorable Hybrid Graphite/Nanodiamond Carbon-Based Material
by Tikhon S. Kurkin, Oleg V. Lebedev, Evgeny K. Golubev, Andrey K. Gatin, Victoria V. Nepomnyashchikh, Valery Yu. Dolmatov and Alexander N. Ozerin
Nanomaterials 2025, 15(1), 56; https://doi.org/10.3390/nano15010056 - 1 Jan 2025
Viewed by 503
Abstract
The results of a comprehensive investigation into the structure and properties of nanodiamond soot (NDS), obtained from the detonation of various explosive precursors (trinitrotoluene, a trinitrotoluene/hexogen mixture, and tetryl), are presented. The colloidal behavior of the NDS particles in different liquid media was [...] Read more.
The results of a comprehensive investigation into the structure and properties of nanodiamond soot (NDS), obtained from the detonation of various explosive precursors (trinitrotoluene, a trinitrotoluene/hexogen mixture, and tetryl), are presented. The colloidal behavior of the NDS particles in different liquid media was studied. The results of the scanning electron microscopy, dynamic light scattering, zeta potential measurements, and laser diffraction analysis suggested a similarity in the morphology of the NDS particle aggregates and agglomerates. The phase composition of the NDS nanoparticles was studied using X-ray diffraction, Raman spectroscopy, electron diffraction, transmission electron microscopy, atomic force microscopy, and scanning tunneling microscopy. The NDS particles were found to comprise both diamond and graphite phases. The ratio of diamond to graphite phase content varied depending on the NDS explosive precursor, while the graphite phase content had a significant impact on the electrical conductivity of NDS. The study of the mechanical and tribological characteristics of polymer nanocomposites, modified with the selected NDS particles, indicated that NDS of various types can serve as a viable set of model nanofillers. Full article
(This article belongs to the Section 2D and Carbon Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>SEM images of the NDS powders: (<b>a</b>,<b>b</b>)—NDS-1; (<b>c</b>,<b>d</b>)—NDS-2; (<b>e</b>,<b>f</b>)—NDS-3.</p>
Full article ">Figure 1 Cont.
<p>SEM images of the NDS powders: (<b>a</b>,<b>b</b>)—NDS-1; (<b>c</b>,<b>d</b>)—NDS-2; (<b>e</b>,<b>f</b>)—NDS-3.</p>
Full article ">Figure 2
<p>Distributions of hydrodynamic diameter of the (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) NDS-1, (<b>b</b>,<b>e</b>,<b>h</b>) NDS-2, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>k</b>) NDS-3 nanoparticles in (<b>a</b>–<b>c</b>) ethanol, (<b>d</b>–<b>f</b>) isopropanol, (<b>g</b>–<b>i</b>) DMSO, and (<b>j</b>,<b>k</b>) acetone, obtained using the DLS technique.</p>
Full article ">Figure 3
<p>Results of zeta potential measurements for (<b>a</b>) NDS-1, (<b>b</b>) NDS-2, and (<b>c</b>) NDS-3 nanoparticles dispersed in different media versus the time between the ultrasonication procedure and the measurement of zeta potential of the NDS aggregates/agglomerates in the dispersions.</p>
Full article ">Figure 4
<p>Distributions of diameter of the NDS particles in a suspension of NDS powders in water, obtained with the LD method for all the investigated NDS powders.</p>
Full article ">Figure 5
<p>TEM images of the NDS powders: (<b>a</b>)—NDS-1; (<b>b</b>)—NDS-2; (<b>c</b>,<b>d</b>)—NDS-3.</p>
Full article ">Figure 5 Cont.
<p>TEM images of the NDS powders: (<b>a</b>)—NDS-1; (<b>b</b>)—NDS-2; (<b>c</b>,<b>d</b>)—NDS-3.</p>
Full article ">Figure 6
<p>XRD patterns (CuK<sub>α</sub>-radiation) for various NDS types in (<b>a</b>) <span class="html-italic">I</span> − 2<span class="html-italic">θ</span> and in (<b>b</b>) <span class="html-italic">I</span>(<span class="html-italic">s</span>)<span class="html-italic">s</span><sup>2</sup><span class="html-italic">ds</span> − <span class="html-italic">s</span> coordinates. The peaks that are colored blue correspond to nanographite, while the peaks that are colored purple correspond to detonation nanodiamonds. Dashed vertical lines in (<b>a</b>) indicate the centers of the peaks.</p>
Full article ">Figure 7
<p>ND weight fraction values in the NDS powders of various types calculated using integration of the corresponding peaks in <span class="html-italic">I</span> − 2<span class="html-italic">θ</span> (“XRD”) and <span class="html-italic">I</span>(<span class="html-italic">s</span>)<span class="html-italic">s</span><sup>2</sup><span class="html-italic">ds</span> − <span class="html-italic">s</span> (“XRD s<sup>2</sup>”) coordinates. Literature data were taken from the publication of Dolmatov et al. [<a href="#B22-nanomaterials-15-00056" class="html-bibr">22</a>].</p>
Full article ">Figure 8
<p>Raman spectroscopy results for NDSs of different types. Dotted lines represent the obtained experimental data, and solid lines are peaks from decomposition of the experimental data. Dashed vertical lines indicate the centers of the peaks.</p>
Full article ">Figure 9
<p>Electron diffraction results for the (<b>a</b>) NDS-1, (<b>b</b>) NDS-2, and (<b>c</b>) NDS-3. The insets in the diffractograms show distributions of normalized intensity along the dashed yellow lines. The yellow arrows point at the scattering peaks, corresponding to the diamond 111 and 022, and graphite 002 planes of carbon atoms.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>c</b>) Topography maps, (<b>d</b>–<b>f</b>) tunnel current distributions, (<b>g</b>–<b>i</b>) overlays of the corresponding tunnel current distributions with the topography maps, and (<b>j</b>–<b>l</b>) voltage-current curves obtained using the STM method for the (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) NDS-1, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) NDS-2, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) NDS-3. The voltage-current curves correspond to the points indicated with the same number and color in the overlayed tunnel current maps (<b>g</b>–<b>i</b>).</p>
Full article ">Figure 11
<p>(<b>a</b>) Schematic image of the apparatus used for measuring specific resistivity of the NDS powders. (<b>b</b>) Conductivity of the NDS powders of various types versus volume fraction of the NDS nanoparticles in the measurement cell.</p>
Full article ">Figure 12
<p>The assumed packaging of the ND (a hexagonal with a triangle inside) and nanographite (a hexagonal with three lines inside) subunits for the (<b>a</b>) NDS-1, (<b>b</b>) NDS-2, and (<b>c</b>) NDS-3 nanoparticles.</p>
Full article ">Figure 13
<p>Dependencies of the friction coefficient on the ND content in the NDS powder (<a href="#nanomaterials-15-00056-t002" class="html-table">Table 2</a>) in nanocomposites based on PP modified with (<b>a</b>) 2.5 and (<b>b</b>) 10 wt.% of NDS of various types. The curves were obtained for different durations (1 and 120 min) of annealing at 205 °C.</p>
Full article ">Figure 14
<p>Dependence of the tensile strength on the homogeneous shear deformation ratio for the composites based on disentangled UHMWPE reactor powder and obtained using the solid-state processing approach without (“UHMWPE”) and with the addition of 10 wt.% NDS of various types.</p>
Full article ">
11 pages, 5221 KiB  
Article
Green Chemical Shear-Thickening Polishing of Monocrystalline Silicon
by Jiancheng Xie, Feng Shi, Shanshan Wang, Xing Peng and Qun Hao
Nanomaterials 2024, 14(23), 1866; https://doi.org/10.3390/nano14231866 - 21 Nov 2024
Viewed by 490
Abstract
A green chemical shear-thickening polishing (GC-STP) method was studied to improve the surface precision and processing efficiency of monocrystalline silicon. A novel green shear-thickening polishing slurry composed of silica nanoparticles, alumina abrasive, sorbitol, plant ash, polyethylene glycol, and deionized water was formulated. The [...] Read more.
A green chemical shear-thickening polishing (GC-STP) method was studied to improve the surface precision and processing efficiency of monocrystalline silicon. A novel green shear-thickening polishing slurry composed of silica nanoparticles, alumina abrasive, sorbitol, plant ash, polyethylene glycol, and deionized water was formulated. The monocrystalline silicon was roughly ground using a diamond polishing slurry and then the GC-STP process. The material removal rate (MRR) during GC-STP was 4.568 μmh−1. The material removal mechanism during the processing of monocrystalline silicon via GC-STP was studied using elemental energy spectroscopy and FTIR spectroscopy. After 4 h of the GC-STP process, the surface roughness (Ra) of the monocrystalline silicon wafer was reduced to 0.278 nm, and an excellent monocrystalline silicon surface quality was obtained. This study shows that GC-STP is a green, efficient, and low-damage polishing method for monocrystalline silicon. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of green chemical shear-thickening polishing slurry (GC-STPS): (<b>a</b>) non-disturbed state; (<b>b</b>) disturbed state.</p>
Full article ">Figure 2
<p>TEM images of (<b>a</b>) nano-silica in GC-STPS; (<b>b</b>) Al<sub>2</sub>O<sub>3</sub> abrasive in GC-STPS.</p>
Full article ">Figure 3
<p>The influence of Al<sub>2</sub>O<sub>3</sub> abrasive diameter on MRR and surface roughness (R<sub>a</sub>) of monocrystalline silicon wafers.</p>
Full article ">Figure 4
<p>SEM image of monocrystalline silicon wafer surface after GC-STP: (<b>a</b>) before polishing; (<b>b</b>) after 2 h of GC-STP; (<b>c</b>) after 4 h of GC-STP.</p>
Full article ">Figure 5
<p>Scanning electron microscope image of monocrystalline silicon face shape accuracy and roughness after GC-STP: (<b>a</b>) before polishing; (<b>b</b>) after 2 h of GC-STP; (<b>c</b>) after 4 h of GC-STP.</p>
Full article ">Figure 6
<p>Energy spectra of elements on the surface of monocrystalline silicon: (<b>a</b>) original surface elements; (<b>b</b>) surface elements of monocrystalline silicon after immersion in GC-STPS.</p>
Full article ">Figure 7
<p>FTIR spectra of sorbitol and monocrystalline silicon surface chelation.</p>
Full article ">Figure 8
<p>Chelating equation for the GC-STP of monocrystalline silicon.</p>
Full article ">Figure 9
<p>Diagrammatic representation of the monocrystalline silicon GC-STP mechanism: (<b>a</b>) chemical softening; (<b>b</b>) complexation reaction in GC-STP; (<b>c</b>) removal of the softened layer by the abrasive; (<b>d</b>) ultra-smooth surface of monocrystalline silicon formed after GC-STP.</p>
Full article ">
13 pages, 5003 KiB  
Article
Effects of Crystalline Diamond Nanoparticles on Silicon Thin Films as an Anode for a Lithium-Ion Battery
by Yonhua Tzeng, Cheng-Ying Jhan, Shi-Hong Sung and Yu-Yang Chiou
Batteries 2024, 10(9), 321; https://doi.org/10.3390/batteries10090321 - 11 Sep 2024
Viewed by 1314
Abstract
Crystalline diamond nanoparticles which are 3.6 nm in size adhering to thin-film silicon results in a hydrophilic silicon surface for uniform wetting by electrolytes and serves as a current spreader for the prevention of a local high-lithium-ion current density. The excellent physical integrity [...] Read more.
Crystalline diamond nanoparticles which are 3.6 nm in size adhering to thin-film silicon results in a hydrophilic silicon surface for uniform wetting by electrolytes and serves as a current spreader for the prevention of a local high-lithium-ion current density. The excellent physical integrity of an anode made of diamond on silicon and the long-life and high-capacity-retention cycling performance are thus achieved for lithium-ion batteries. A specific capacity of 1860 mAh/g(si) was retained after 200 cycles of discharge/charge at an areal current density of 0.2 mA/cm2. This is compared to 1626 mAh/g(si) for a thin-film-silicon anode without the additive of diamond nanoparticles. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the thin-film anodes showing the effects of nano-diamonds (NDs) on the uniform lithiation of the anode made of (<b>top</b>) Si and (<b>bottom</b>) Si/D/Si/D. Red arrows represent the transportation of lithium ions. Blue diamond shape represents nano-diamond (ND) particles. The black area represents SEI layers. Samples without diamond layers exhibited structural instability during cycling, leading to drastic changes in local and overall thickness of the anode.</p>
Full article ">Figure 2
<p>SEM cross-sectional images of (<b>a</b>) Si anode and (<b>b</b>) Si/D/Si/D anode; SEM top-view images of the (<b>c</b>) Si anode and (<b>d</b>) Si/D/Si/D anode.</p>
Full article ">Figure 3
<p>Raman spectra of a Si anode and a Si/D/Si/D anode.</p>
Full article ">Figure 4
<p>CV curves of (<b>a</b>) Si anode and (<b>b</b>) Si/D/Si/D anode at a scan rate of 0.05 mV/s. Charge–discharge curves of several cycle numbers for the (<b>c</b>) Si anode and (<b>d</b>) Si/D/Si/D anode.</p>
Full article ">Figure 5
<p>(<b>a</b>) Cycle performance of Si thin-film anode and Si/D/Si/D thin-film anode during the first three cycles of charge and discharge at a current density of 0.02 mA/cm<sup>2</sup>, and for subsequent cycles at a current density of 0.2 mA/cm<sup>2</sup>. (<b>b</b>) C-Rate performance of Si anode and Si/D/Si/D anode.</p>
Full article ">Figure 6
<p>Optical microscope images of (<b>a</b>) Si anode and (<b>b</b>) Si/D/Si/D anode after 200 cycles. (<b>c</b>) Si anode and (<b>d</b>) Si/D/Si/D anode after 80 C-rate testing cycles.</p>
Full article ">Figure 7
<p>SEM images of (<b>a</b>,<b>b</b>) the surface of Si anode after 200 cycles at different magnification, (<b>c</b>) a cross-section of a Si anode; SEM images of (<b>d</b>,<b>e</b>) the surface of a Si/D/Si/D anode after 200 cycles at different magnification, (<b>f</b>) a cross-section of a Si/D/Si/D anode; EDS analysis of (<b>g</b>) a silicon anode and (<b>h</b>) a Si/D/Si/D anode after 200 cycles.</p>
Full article ">Figure 8
<p>XPS (<b>a</b>) C 1s spectra of Si anode and Si/D/Si/D anode; (<b>b</b>) Si 2p spectra of Si anode and Si/D/Si/D anode, SiO<sub>2</sub> has two bands shown in red color; (<b>c</b>) P 2p spectra of Si anode and Si/D/Si/D anode; and (<b>d</b>) F 1s spectra of Si anode and Si/D/Si/D anode after 200 cycles of discharge and charge operation.</p>
Full article ">
26 pages, 8953 KiB  
Article
Silver and Carbon Nanomaterials/Nanocomplexes as Safe and Effective ACE2-S Binding Blockers on Human Skin Cell Lines
by Anna Hotowy, Barbara Strojny-Cieślak, Agnieszka Ostrowska, Marlena Zielińska-Górska, Marta Kutwin, Mateusz Wierzbicki, Malwina Sosnowska, Sławomir Jaworski, André Chwalibóg, Ireneusz Kotela and Ewa Sawosz Chwalibóg
Molecules 2024, 29(15), 3581; https://doi.org/10.3390/molecules29153581 - 29 Jul 2024
Viewed by 1020
Abstract
(1) Background: Angiotensin-converting enzyme 2 (ACE2) is a crucial functional receptor of the SARS-CoV-2 virus. Although the scale of infections is no longer at pandemic levels, there are still fatal cases. The potential of the virus to infect the skin raises questions about [...] Read more.
(1) Background: Angiotensin-converting enzyme 2 (ACE2) is a crucial functional receptor of the SARS-CoV-2 virus. Although the scale of infections is no longer at pandemic levels, there are still fatal cases. The potential of the virus to infect the skin raises questions about new preventive measures. In the context of anti-SARS-CoV-2 applications, the interactions of antimicrobial nanomaterials (silver, Ag; diamond, D; graphene oxide, GO and their complexes) were examined to assess their ability to affect whether ACE2 binds with the virus. (2) Methods: ACE2 inhibition competitive tests and in vitro treatments of primary human adult epidermal keratinocytes (HEKa) and primary human adult dermal fibroblasts (HDFa) were performed to assess the blocking capacity of nanomaterials/nanocomplexes and their toxicity to cells. (3) Results: The nanocomplexes exerted a synergistic effect compared to individual nanomaterials. HEKa cells were more sensitive than HDFa cells to Ag treatments and high concentrations of GO. Cytotoxic effects were not observed with D. In the complexes, both carbonic nanomaterials had a soothing effect against Ag. (4) Conclusions: The Ag5D10 and Ag5GO10 nanocomplexes seem to be most effective and safe for skin applications to combat SARS-CoV-2 infection by blocking ACE2-S binding. These nanocomplexes should be evaluated through prolonged in vivo exposure. The expected low specificity enables wider applications. Full article
(This article belongs to the Special Issue 2D Nanosheets and Their Nanohybrids)
Show Figures

Figure 1

Figure 1
<p>TEM visualization of nanomaterials and their complexes. D—diamond nanoparticles, scale bar reflects 100 nm; GO—graphene oxide, scale bar reflects 500 nm; DAg—nanocomplex of diamond and silver nanoparticles, scale bar reflects 100 nm; GOAg—nanocomplex of graphene oxide and silver nanoparticles, scale bar reflects 1 µm; Ag—silver nanoparticles, scale bar reflects 200 nm. GOAg in the blue and yellow frame—scale bars reflect 200 nm; GOAg in the green frame—scale bar reflects 100 nm.</p>
Full article ">Figure 2
<p>ACE2-S binding in the presence of the specified nanomaterials. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–d above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups.</p>
Full article ">Figure 3
<p>ACE2-S binding in the presence of specified nanomaterials. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–c above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 4
<p>Viability of HEKa cells treated with different concentrations of carbon nanostructures and silver nanoparticles. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–h above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 5
<p>Viability of HDFa treated with different concentrations of carbon nanostructures and silver nanoparticles. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–e above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 6
<p>HEKa morphology in a culture treated with different concentrations of carbon nanostructures and silver nanoparticles. The red frame indicates major morphological changes in cells treated with the highest concentration of nanomaterials. Ag—10 µg/mL, D—100 µg/mL, GO—100 µg/mL. Scale bar 200 µm.</p>
Full article ">Figure 7
<p>HDFa morphology in a culture treated with different concentrations of carbon nanostructures and silver nanoparticles. The red frame indicates major morphological changes in cells treated with the highest concentration of nanomaterials. Ag—10 µg/mL, D—100 µg/mL, GO—100 µg/mL. Scale bar 200 µm.</p>
Full article ">Figure 8
<p>Viability of HEKa cells treated with (<b>a</b>) diamond and (<b>b</b>) graphene oxide complexes with silver nanoparticles of different concentrations. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–g above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 9
<p>Viability of HEKa cells treated with diamond and graphene oxide complexes with silver nanoparticles at a concentration of (<b>a</b>) 2 µg/mL and (<b>b</b>) 5 µg/mL. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–e above the treatment bars—the same letters indicate no significant differences between the treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 10
<p>Viability of HDFa treated with (<b>a</b>) diamond and (<b>b</b>) graphene oxide complexes with silver nanoparticles of different concentrations. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–f above the treatment bars—the same letters indicate no significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 11
<p>Viability of HDFa treated with diamond and graphene oxide complexes with silver nanoparticles at concentrations of (<b>a</b>) 2 µg/mL and (<b>b</b>) 5 µg/mL. Results are presented as % of control (means ± SD, n = 6). <span class="html-italic">p</span> ≤ 0.05 was considered a statistically significant difference. Letters a–e above the treatment bars—different letters indicate significant differences between treatments, while different letters above treatment bars indicate the presence of statistically significant differences between specific groups. Nanomaterial concentrations refer to micrograms per millilitre.</p>
Full article ">Figure 12
<p>HEKa (<b>a</b>) and HDFa (<b>b</b>) cell morphology in a culture treated with carbon nanostructures, silver nanoparticles, and their complexes (blue frame). Red arrowheads—shrunken cells; yellow arrowheads—vacuolized cytoplasm; green arrowheads—shortened filopodia; blue arrowheads—nuclear chromatin accumulated near the nuclear membrane; purple arrowheads—apoptotic bodies visible as protrusions from the plasma membrane.</p>
Full article ">
21 pages, 21058 KiB  
Article
Tribocatalytic Reaction Enabled by TiO2 Nanoparticle for MoDTC-Derived Tribofilm Formation at ta-C/Steel Contact
by Daiki Matsukawa, Jae-Hyeok Park, Woo-Young Lee, Takayuki Tokoroyama, Jae-Il Kim, Ryoichi Ichino and Noritsugu Umehara
Coatings 2024, 14(6), 773; https://doi.org/10.3390/coatings14060773 - 19 Jun 2024
Viewed by 1549
Abstract
Tribochemically produced triboproducts are becoming increasingly important in tribosystems and serve to improve system performance by preventing friction or wear. Diamond-like carbon (DLC) is chemically stable, which features a trade-off with tribological pros and cons. Chemically stable DLC is thermally stable and suppresses [...] Read more.
Tribochemically produced triboproducts are becoming increasingly important in tribosystems and serve to improve system performance by preventing friction or wear. Diamond-like carbon (DLC) is chemically stable, which features a trade-off with tribological pros and cons. Chemically stable DLC is thermally stable and suppresses surface damage in a high-temperature operating environment; however, it causes a detrimental effect that hinders the formation of a competent tribofilm. In this study, we dispersed highly reactive TiO2 nanoparticles (TDONPs) in molybdenum dithiocarbamate (MoDTC)-containing lubricant for adhering triboproducts on the DLC surface. In addition, TDONPs contributed to the decomposition of triboproducts by promoting the decomposition of MoDTC through its catalytic role. Rutile TDONPs were more helpful in reducing friction than anatase TDONPs and improved the friction performance by up to ~100%. Full article
(This article belongs to the Special Issue Advanced Tribological Coatings: Fabrication and Application)
Show Figures

Figure 1

Figure 1
<p>Schematic of hybrid coating system equipped with an anode-layer ion source, filtered cathodic vacuum arc source, and unbalanced magnetron sputter source.</p>
Full article ">Figure 2
<p>Images after tribotest of lubricant composed of PAO4 base oil with 2 wt.% of (<b>a</b>) TDONPs and (<b>b</b>) OA-modified TDONPs and (<b>c</b>) 2.5 wt.% of OA-modified TDONPs.</p>
Full article ">Figure 3
<p>(<b>a</b>) Friction curves of ta-C disk/steel ball tribopair under PAO4 mixed with MoDTC 700 ppm and various addition amounts of TDONPs: (<b>a</b>) a-TDONPs and (<b>b</b>) r-TDONPs.</p>
Full article ">Figure 4
<p>(<b>a</b>) Average friction coefficient for steady-state (240–250 m) and (<b>b</b>) specific wear rate of ta-C disk under PAO4 mixed with MoDTC 700 ppm and various addition amounts of TDONPs.</p>
Full article ">Figure 5
<p>Optical images of wear track on ta-C disk and wear scar of steel ball after friction test under PAO4 mixed with MoDTC 700 ppm and TDONPs: (<b>a</b>) a-TDONPs and (<b>b</b>) r-TDONPs.</p>
Full article ">Figure 6
<p>Images of SEM and EDS mapping on (<b>a</b>) ta-C disk and (<b>b</b>) steel ball. w/o NPs, a—2.0 wt.%, r—2.0 wt.% indicate worn surfaces lubricated under without NPs, and with 2 wt.% of a-TDONPs and r-TDONPs.</p>
Full article ">Figure 7
<p>Raman spectra of tribofilm formed on the worn surface of steel ball: (<b>a</b>) without NPs, and with (<b>b</b>) a-TDONPs 2 wt.% or (<b>c</b>) r-TDONPs 2 wt.%. Red-colored circles indicate acquired points.</p>
Full article ">Figure 8
<p>Raman spectra of tribofilm formed on worn surface of ta-C disk. Red-colored circles indicate acquired points.</p>
Full article ">Figure 9
<p>Respective XPS spectra for (<b>a</b>) Ti 2<span class="html-italic">p</span>, (<b>b</b>) Mo 3<span class="html-italic">d</span>, and (<b>c</b>) S 2<span class="html-italic">p</span> of worn surface on ta-C disks under PAO4 oil with 2 wt.% r-TDONPs and 700 ppm MoDTC. Depth profiling XPS analysis for (<b>d</b>) Ti 2<span class="html-italic">p</span>, (<b>e</b>) Mo 3<span class="html-italic">d</span>, and (<b>f</b>) S 2<span class="html-italic">p</span> of tribofilm formed on ta-C slid under 2 wt.% r-TDONPs containing MoDTC lubricant.</p>
Full article ">Figure 10
<p>Depth profile of atomic concentration of tribofilms on ta-C disks lubricated without (<b>a</b>) TDONPs, and with 2 wt.% of (<b>b</b>) r-TDONPs and (<b>c</b>) a-TDONPs.</p>
Full article ">Figure 11
<p>Depth profile of atomic concentration of tribofilms on steel balls lubricated (<b>a</b>) without TDONPs, and with 2 wt.% of (<b>b</b>) r-TDONPs and (<b>c</b>) a-TDONPs.</p>
Full article ">Figure 12
<p>Area percent in (<b>a</b>–<b>c</b>) Mo 3<span class="html-italic">d</span> and (<b>d</b>,<b>e</b>) Ti 2<span class="html-italic">p</span> of tribofilms on ta-C disks: (<b>a</b>) without TDONPs, (<b>b</b>,<b>d</b>) with r-TDONPs 2 wt.%, and (<b>c</b>,<b>e</b>) with a-TDONPs 2 wt.%.</p>
Full article ">Figure 13
<p>Area percent in (<b>a</b>–<b>c</b>) Mo 3<span class="html-italic">d</span> and (<b>d</b>,<b>e</b>) Ti 2<span class="html-italic">p</span> of tribofilms on steel balls: (<b>a</b>) without TDONPs, (<b>b</b>,<b>d</b>) with r-TDONPs 2 wt.%, and (<b>c</b>,<b>e</b>) with a-TDONPs 2 wt.%.</p>
Full article ">Figure 14
<p>Atomic ratio of (<b>a</b>,<b>d</b>) S/Mo, (<b>b</b>,<b>e</b>) area ratio of Mo<sup>0</sup> and Mo<sup>2+</sup> in Mo 3<span class="html-italic">d</span>, and (<b>c</b>,<b>f</b>) Ti<sup>4+</sup> in Ti 2<span class="html-italic">p</span> in tribofilm: (<b>a</b>–<b>c</b>) ta-C disks and (<b>d</b>–<b>f</b>) steel balls.</p>
Full article ">Figure 15
<p>XPS spectra for (<b>a</b>) survey, (<b>b</b>) S 2<span class="html-italic">p</span>, and (<b>c</b>) Mo 3<span class="html-italic">d</span> of worn surface on ta-C disks lubricated without TDONPs and with a-TDONPs and r-TDONPs.</p>
Full article ">Figure 16
<p>(<b>a</b>) Atomic concentration of worn surface on ta-C disks and area percent in (<b>b</b>) Mo 3<span class="html-italic">d</span>, (<b>c</b>) Ti 2<span class="html-italic">p</span>, and (<b>d</b>) S 2<span class="html-italic">p</span> lubricated without TDONPs, and with 2 wt.% of r-TDONPs and a-TDONPs. (<b>e</b>) Atomic ratio of S/Mo and S/O, and bonding ratio of Mo<sup>4+</sup> of worn ta-C disks.</p>
Full article ">Figure 17
<p>(<b>a</b>) Atomic concentration of worn surface on steel balls and area percent in (<b>b</b>) Mo 3<span class="html-italic">d</span>, (<b>c</b>) Ti 2<span class="html-italic">p</span>, and (<b>d</b>) S 2<span class="html-italic">p</span> lubricated without TDONPs, and with 2 wt.% of r-TDONPs and a-TDONPs. (<b>e</b>) Atomic ratio of S/Mo and S/O, and bonding ratio of Mo<sup>4+</sup> of worn steel balls.</p>
Full article ">Figure 18
<p>Relationship between friction coefficient and atomic ratio of Mo/S and O/S and bonding ratio of Mo<sup>4+</sup>: (<b>a</b>) ta-C disk and (<b>b</b>) steel ball.</p>
Full article ">Figure 19
<p>(<b>a</b>) Friction curves, (<b>b</b>) average friction coefficient, and (<b>c</b>) specific wear rate of ta-C under PAO4 mixed with MoDTC 700 ppm and 2 wt.% a-TDONPs.</p>
Full article ">Figure 20
<p>(<b>a</b>) Friction curves and (<b>b</b>) average friction coefficient for steady-state ta-C disk/steel ball tribopair under PAO4 with various addition amounts of a-TDONPs.</p>
Full article ">Figure 21
<p>(<b>a</b>) Optical images and (<b>b</b>,<b>c</b>) EDS mapping images of wear scar on steel ball lubricated under PAO with various addition amounts of a-TDONPs.</p>
Full article ">Figure 22
<p>Ti/C atomic ratio measured using AES of tribofilm in depth direction on a steel ball and ta-C disk.</p>
Full article ">Figure 23
<p>Schematic of (<b>a</b>) the expected tribofilm structure and (<b>b</b>) degradation of MoDTC with TDONPs.</p>
Full article ">
19 pages, 10295 KiB  
Article
Production of Cu/Diamond Composite Coatings and Their Selected Properties
by Grzegorz Cieślak, Marta Gostomska, Adrian Dąbrowski, Katarzyna Skroban, Tinatin Ciciszwili-Wyspiańska, Edyta Wojda, Anna Mazurek, Michał Głowacki, Michał Baranowski, Anna Gajewska-Midziałek and Maria Trzaska
Materials 2024, 17(12), 2803; https://doi.org/10.3390/ma17122803 - 8 Jun 2024
Cited by 1 | Viewed by 1184
Abstract
This article presents Cu/diamond composite coatings produced by electrochemical reduction on steel substrates and a comparison of these coatings with a copper coating without diamond nanoparticles (<10 nm). Deposition was carried out using multicomponent electrolyte solutions at a current density of 3 A/dm [...] Read more.
This article presents Cu/diamond composite coatings produced by electrochemical reduction on steel substrates and a comparison of these coatings with a copper coating without diamond nanoparticles (<10 nm). Deposition was carried out using multicomponent electrolyte solutions at a current density of 3 A/dm2 and magnetic stirring speed of 100 rpm. Composite coatings were deposited from baths with different diamond concentrations (4, 6, 8, 10 g/dm3). This study presents the surface morphology and structure of the produced coatings. The surface roughness, coating thickness (XRF), mechanical properties (DSI), and adhesion of coatings to substrates (scratch tests) were also characterized. The coatings were also tested to assess their solderability, including their spreadability, wettability of the solder, durability of solder-coating bonds, and a microstructure study. Full article
Show Figures

Figure 1

Figure 1
<p>Samples (<b>a</b>) for spreadability and microstructure tests (substrate-coating-solder), (<b>b</b>) for adhesion test. (unit: mm).</p>
Full article ">Figure 1 Cont.
<p>Samples (<b>a</b>) for spreadability and microstructure tests (substrate-coating-solder), (<b>b</b>) for adhesion test. (unit: mm).</p>
Full article ">Figure 2
<p>Surface morphology and XRD pattern of the used diamond.</p>
Full article ">Figure 3
<p>Surface morphology and XRD patterns of produced coatings: (<b>a</b>) Cu; (<b>b</b>) Cu/diamond (4); (<b>c</b>) Cu/diamond (6); (<b>d</b>) Cu/diamond (8); (<b>e</b>) Cu/diamond (10); ICDD 00-004-0836.</p>
Full article ">Figure 3 Cont.
<p>Surface morphology and XRD patterns of produced coatings: (<b>a</b>) Cu; (<b>b</b>) Cu/diamond (4); (<b>c</b>) Cu/diamond (6); (<b>d</b>) Cu/diamond (8); (<b>e</b>) Cu/diamond (10); ICDD 00-004-0836.</p>
Full article ">Figure 4
<p>Dependence of load on the depth of penetration of the indenter into the material of the tested Cu and Cu/diamond coatings.</p>
Full article ">Figure 5
<p>Images of scratches after scratch testing: (<b>a</b>) Cu; (<b>b</b>) Cu/diamond (4); (<b>c</b>) Cu/diamond (6); (<b>d</b>) Cu/diamond (8); (<b>e</b>) Cu/diamond (10).</p>
Full article ">Figure 6
<p>Frictional force and friction coefficient during the scratch test of produced coatings: (<b>a</b>) Cu; (<b>b</b>) Cu/diamond (4); (<b>c</b>) Cu/diamond (6); (<b>d</b>) Cu/diamond (8); (<b>e</b>) Cu/diamond (10).</p>
Full article ">Figure 6 Cont.
<p>Frictional force and friction coefficient during the scratch test of produced coatings: (<b>a</b>) Cu; (<b>b</b>) Cu/diamond (4); (<b>c</b>) Cu/diamond (6); (<b>d</b>) Cu/diamond (8); (<b>e</b>) Cu/diamond (10).</p>
Full article ">Figure 7
<p>Solder spreadability on Cu coating: (<b>a</b>) SnCu1 solder alloy, (<b>b</b>) SAC305 solder alloy, (<b>c</b>) Sn60Pb40 solder alloy.</p>
Full article ">Figure 8
<p>Solder spreadability on Cu/diamond (10) coating: (<b>a</b>) SnCu1 solder alloy, (<b>b</b>) SAC305 solder alloy, (<b>c</b>) Sn60Pb40 solder alloy.</p>
Full article ">Figure 9
<p>Selected contact angle measurements: (<b>a</b>) Sn60Pb40 solder alloy Cu/diamond (10) coating; (<b>b</b>) SnCu1 solder alloy Cu/diamond (10) coating.</p>
Full article ">Figure 9 Cont.
<p>Selected contact angle measurements: (<b>a</b>) Sn60Pb40 solder alloy Cu/diamond (10) coating; (<b>b</b>) SnCu1 solder alloy Cu/diamond (10) coating.</p>
Full article ">Figure 10
<p>Linear distribution of elements in the contact zone between solder and Cu coating: (<b>a</b>) SnCu1; (<b>b</b>) SAC305; (<b>c</b>) Sn60Pb40.</p>
Full article ">Figure 11
<p>Linear distribution of elements in the contact zone between solder and Cu/diamond (10) coating: (<b>a</b>) SnCu1; (<b>b</b>) SAC305; (<b>c</b>) Sn60Pb40.</p>
Full article ">Figure 12
<p>Results of the adhesion test.</p>
Full article ">
15 pages, 9244 KiB  
Article
XMEA: A New Hybrid Diamond Multielectrode Array for the In Situ Assessment of the Radiation Dose Enhancement by Nanoparticles
by Patricia Nicolucci, Guilherme Gambaro, Kyssylla Monnyelle Araujo Silva, Iara Souza Lima, Oswaldo Baffa and Alberto Pasquarelli
Sensors 2024, 24(8), 2409; https://doi.org/10.3390/s24082409 - 10 Apr 2024
Viewed by 1085
Abstract
This work presents a novel multielectrode array (MEA) to quantitatively assess the dose enhancement factor (DEF) produced in a medium by embedded nanoparticles. The MEA has 16 nanocrystalline diamond electrodes (in a cell-culture well), and a single-crystal diamond divided into four quadrants for [...] Read more.
This work presents a novel multielectrode array (MEA) to quantitatively assess the dose enhancement factor (DEF) produced in a medium by embedded nanoparticles. The MEA has 16 nanocrystalline diamond electrodes (in a cell-culture well), and a single-crystal diamond divided into four quadrants for X-ray dosimetry. DEF was assessed in water solutions with up to a 1000 µg/mL concentration of silver, platinum, and gold nanoparticles. The X-ray detectors showed a linear response to radiation dose (r2 ≥ 0.9999). Overall, platinum and gold nanoparticles produced a dose enhancement in the medium (maximum of 1.9 and 3.1, respectively), while silver nanoparticles produced a shielding effect (maximum of 37%), lowering the dose in the medium. This work shows that the novel MEA can be a useful tool in the quantitative assessment of radiation dose enhancement due to nanoparticles. Together with its suitability for cells’ exocytosis studies, it proves to be a highly versatile device for several applications. Full article
(This article belongs to the Section Biosensors)
Show Figures

Figure 1

Figure 1
<p>Structures and assembly of the XMEA: (<b>a</b>) Layout of the 16 Ch MEA. Channels 2, 7, 10, and 15 are not connected, leaving the connections available for the X-ray photodiodes; (<b>b</b>) magnification of the sensing area with 20 µm large µ-electrodes; (<b>c</b>) layout of the 4-quadrant X-ray detector consisting of p-i-m-photodiodes; (<b>d</b>) top view of the XMEA stack. The back aluminum contact has an opening for allowing the inspection of the MEA sensing area with an inverted microscope, thanks to the semi-transparency of the XMEA; (<b>e</b>) cross-section of the fully assembled XMEA. The glue between the intrinsic diamond layer (i-NCD) and the glass carrier is not shown.</p>
Full article ">Figure 2
<p>UV–vis spectra of colloidal solutions of different nanoparticles: (<b>a</b>) AgNPs, (<b>b</b>) AuNPs, and (<b>c</b>) PtNPs. The UV–vis spectra provide essential optical information about the nanoparticles, revealing characteristic absorption peaks that correspond to their unique electronic transition. (<b>d</b>) The average hydrodynamic size distribution of the metallic nanoparticles obtained using the dynamic light-scattering technique.</p>
Full article ">Figure 3
<p>Transmission electron microscopy images (<b>left</b>) and size (diameter) distribution (<b>right</b>) of (<b>a</b>) AgNPs, (<b>b</b>) PtNPs, and (<b>c</b>) AuNPs. The images provide detailed visualizations of the spherical morphology and structure of each nanoparticle type, highlighting their distinctive characteristics.</p>
Full article ">Figure 4
<p>Typical signal of the radiation detectors of the d-MEA device when exposed to X-rays.</p>
Full article ">Figure 5
<p>Response of the d-MEA as a function of air-kerma produced by X-rays.</p>
Full article ">Figure 6
<p>Dose enhancement factors obtained for the d-MEA device for AgNP, PtNP, and AuNP as functions of nanoparticle concentration.</p>
Full article ">Figure 7
<p>Response enhancement factors (REF) for the X-ray detectors of XMEA for AgNP, PtNP, and AuNP as a function of nanoparticle concentration.</p>
Full article ">
20 pages, 11077 KiB  
Article
Experimental and Theoretical Investigation on Heat Transfer Enhancement in Micro Scale Using Helical Connectors
by Malyne Abraham, Zachary Abboud, Gabriel Herrera Arriaga, Kendall Tom, Samuel Austin and Saeid Vafaei
Materials 2024, 17(5), 1067; https://doi.org/10.3390/ma17051067 - 26 Feb 2024
Cited by 1 | Viewed by 1134
Abstract
Microscale electronics have become increasingly more powerful, requiring more efficient cooling systems to manage the higher thermal loads. To meet this need, current research has been focused on overcoming the inefficiencies present in typical thermal management systems due to low Reynolds numbers within [...] Read more.
Microscale electronics have become increasingly more powerful, requiring more efficient cooling systems to manage the higher thermal loads. To meet this need, current research has been focused on overcoming the inefficiencies present in typical thermal management systems due to low Reynolds numbers within microchannels and poor physical properties of the working fluids. For the first time, this research investigated the effects of a connector with helical geometry on the heat transfer coefficient at low Reynolds numbers. The introduction of a helical connector at the inlet of a microchannel has been experimentally tested and results have shown that this approach to flow augmentation has a great potential to increase the heat transfer capabilities of the working fluid, even at low Reynolds numbers. In general, a helical connector can act as a stabilizer or a mixer, based on the characteristics of the connector for the given conditions. When the helical connector acts as a mixer, secondary flows develop that increase the random motion of molecules and possible nanoparticles, leading to an enhancement in the heat transfer coefficient in the microchannel. Otherwise, the heat transfer coefficient decreases. It is widely known that introducing nanoparticles into the working fluids has the potential to increase the thermal conductivity of the base fluid, positively impacting the heat transfer coefficient; however, viscosity also tends to increase, reducing the random motion of molecules and ultimately reducing the heat transfer capabilities of the working fluid. Therefore, optimizing the effects of nanoparticles characteristics while reducing viscous effects is essential. In this study, deionized water and deionized water–diamond nanofluid at 0.1 wt% were tested in a two-microchannel system fitted with a helical connector in between. It was found that the helical connector can make a great heat transfer coefficient enhancement in low Reynolds numbers when characteristics of geometry are optimized for given conditions. Full article
(This article belongs to the Section Advanced Nanomaterials and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Schematic of two-microchannel experimental setup with connector in between.</p>
Full article ">Figure 2
<p>Enlarged, dimensioned schematic of a stainless-steel microchannel showing how temperature distribution measurements were taken at the inlet and outlet and along the channel.</p>
Full article ">Figure 3
<p>Dimensioned diagram of the nozzle-shaped connector used for result comparison.</p>
Full article ">Figure 4
<p>Schematic of a helical connector with labeled geometric parameters.</p>
Full article ">Figure 5
<p>Streamlines in a helical connector when <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>40</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>4</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>. Working fluid was deionized water, and Re = 400.</p>
Full article ">Figure 6
<p>Streamlines in a helical connector when <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>40</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>4</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>. Working fluid was deionized water, and Re = 400.</p>
Full article ">Figure 7
<p>Streamlines in a helical connector when <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>60</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, and <span class="html-italic">p</span> = 4 mm. Working fluid was deionized water, and Re = 400.</p>
Full article ">Figure 8
<p>Velocity vectors, streamlines, and velocity contours of a helical connector when <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>40</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>4</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>. Working fluid was deionized water, and Re = 400.</p>
Full article ">Figure 9
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>. Working fluid was deionized water, Re = 300, <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>40</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>10</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.5</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>4</mn> <mo> </mo> <mi>mm</mi> </mrow> </semantics></math>. The red dotted line displays the variation of heat transfer coefficient in each microchannel and error bars have been included displaying the standard deviation of heat transfer coefficient at each point measured.</p>
Full article ">Figure 10
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> for connector (I) (blue) and connector (II) (black). Working fluid was deionized water, (<b>a</b>) Re = 400 and (<b>b</b>) Re = 100.</p>
Full article ">Figure 11
<p>Heat transfer coefficient vs. non-dimensional location <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>, for connector (I) (blue) and connector (II) (red). Working fluid was deionized water, (<b>a</b>) Re = 400 and (<b>b</b>) Re = 100.</p>
Full article ">Figure 12
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>, for connector (III) (red) and connector (IV) (blue). Working fluid was deionized water, (<b>a</b>) Re = 400 and (<b>b</b>) Re = 100.</p>
Full article ">Figure 13
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>, with helical connector (I) in between. Working fluid was deionized water, Re = 12.5–500.</p>
Full article ">Figure 14
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>, in connector (I). Working fluid was deionized water (black) and deionized water–nanodiamond nanofluid (red) with mass concentration of 0.1 wt%, Re = 12.5–400.</p>
Full article ">Figure 15
<p>Heat transfer coefficient vs. non-dimensional location, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> for connector (I) (blue) and nozzle-shaped connector (red) [<a href="#B25-materials-17-01067" class="html-bibr">25</a>]. Working fluid was deionized water, Re = 400.4.</p>
Full article ">
12 pages, 3502 KiB  
Article
The Role of Diamonds Dispersed in Ferronematic Liquid Crystals on Structural Properties
by Peter Bury, Marek Veveričík, František Černobila, Natália Tomašovičová, Veronika Lacková, Katarína Kónyová, Ivo Šafařík, Viktor Petrenko, Oleksandr Tomchuk, Milan Timko and Peter Kopčanský
Crystals 2024, 14(3), 202; https://doi.org/10.3390/cryst14030202 - 20 Feb 2024
Viewed by 1294
Abstract
A study of the role of diamond nanoparticles on 5CB liquid crystal composites with Fe3O4 nanoparticles is presented. Composite ferronematic systems based on the nematic liquid crystal 5CB doped with Fe3O4 magnetic nanoparticles and additionally bound to [...] Read more.
A study of the role of diamond nanoparticles on 5CB liquid crystal composites with Fe3O4 nanoparticles is presented. Composite ferronematic systems based on the nematic liquid crystal 5CB doped with Fe3O4 magnetic nanoparticles and additionally bound to diamond nanoparticles (DNPs), of a volume concentration of 3.2 mg/mL, 1.6 mg/mL and 0.32 mg/mL, were investigated using both magneto-optical effect and surface acoustic waves (SAWs) to study the role of diamond nanoparticles on the structural properties of ferronematic liquid crystals. The responses of light transmission and SAW attenuation to an external magnetic field were investigated experimentally under a linearly increasing and decreasing magnetic field, respectively. Investigations of the phase transition temperature shift of individual composites were also performed. The experimental results highlighted a decrease in the threshold field in the ferronematic LC composites compared to the pure 5CB as well as its further decrease after mixing Fe3O4 with diamond powder. Concerning the transition temperature, its increase with an increase in the volume fraction of both kinds of nanoparticles was registered. The role of diamond nanoparticles in the structural changes and the large residual light transition and/or attenuation (memory effect) were also observed. The presented results confirmed the potential of diamond nanoparticles in nematic composites to modify their properties which could lead to final applications. Full article
(This article belongs to the Section Liquid Crystals)
Show Figures

Figure 1

Figure 1
<p>Bright-field TEM images of studied diamond-bonded ferroparticles.</p>
Full article ">Figure 2
<p>Dependence of light transmission on magnetic field for 5CB composites with Fe<sub>3</sub>O<sub>4</sub> nanoparticles of concentrations 0.32, 1.60 and 3.20 mg/mL (<b>a</b>) and Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs (<b>b</b>), both including pure 5CB, measured in increasing regimes. The measurements were conducted at 25 °C.</p>
Full article ">Figure 3
<p>Dependence of light transmission on magnetic field for 5CB composites with Fe<sub>3</sub>O<sub>4</sub> nanoparticles of concentrations 0.32, 1.60 and 3.20 mg/mL (<b>a</b>) and Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs (<b>b</b>), both including pure 5CB, measured in increasing (full lines) and decreasing regimes (dotted lines). The measurements were conducted at 25 °C.</p>
Full article ">Figure 4
<p>Comparison of light transmission dependences on magnetic field for 5CB composites doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles and Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs, both of concentration 3.20 mg/mL, including pure 5CB measured in increasing (full lines) and decreasing regimes (dotted lines). The measurements were conducted at 25 °C.</p>
Full article ">Figure 5
<p>SAW attenuation response of 5CB composites doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles of different concentrations (0.32, 1.60 and 3.20 mg/mL) (<b>a</b>) and Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs (<b>b</b>) both including pure 5CB, measured on magnetic field in increasing regimes. The measurements were conducted at 25 °C.</p>
Full article ">Figure 6
<p>SAW attenuation response of 5CB composites doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles of different concentrations (0.32, 1.60 and 3.20 mg/mL) (<b>a</b>) and Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs (<b>b</b>), both including pure 5CB, measured in increasing (full lines) and decreasing regimes (dotted lines). The measurements were conducted at 25 °C.</p>
Full article ">Figure 7
<p>SAW attenuation dependences on magnetic field for 5CB composite doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs of concentration 3.20 mg/mL measured in increasing (full lines) and decreasing regimes (dotted lines) three times one after another in 25 min. intervals. The measurements were conducted at 25 °C.</p>
Full article ">Figure 8
<p>SAW attenuation response of 5CB composites doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs of concentrations 0.32, 1.60 and 3.20 mg/mL on magnetic field measured for <b><span class="html-italic">B</span></b> parallel with LC cells in increasing (full lines) and decreasing (dotted lines) regime. The measurements were conducted at 25 °C.</p>
Full article ">Figure 9
<p>Dependences of SAW attenuation response changes on temperature for 5CB composites doped with Fe<sub>3</sub>O<sub>4</sub> nanoparticles and with Fe<sub>3</sub>O<sub>4</sub> nanoparticles bound to DNPs of concentrations 0.32, 1.60 and 3.20 mg/mL measured near the transition temperature.</p>
Full article ">
38 pages, 8233 KiB  
Review
Powders of Diamond Nanoparticles as a Promising Material for Reflectors of Very Cold and Cold Neutrons
by Egor Lychagin, Marc Dubois and Valery Nesvizhevsky
Nanomaterials 2024, 14(4), 387; https://doi.org/10.3390/nano14040387 - 19 Feb 2024
Cited by 1 | Viewed by 1832
Abstract
More than 15 years ago, the study of nanodiamond (ND) powders as a material for designing reflectors of very cold neutrons (VCNs) and cold neutrons (CNs) began. Such reflectors can significantly increase the efficiency of using such neutrons and expand the scope of [...] Read more.
More than 15 years ago, the study of nanodiamond (ND) powders as a material for designing reflectors of very cold neutrons (VCNs) and cold neutrons (CNs) began. Such reflectors can significantly increase the efficiency of using such neutrons and expand the scope of their application for solving applied and fundamental problems. This review considers the principle of operation of VCN and CN reflectors based on ND powders and their advantages. Information is presented on the performed experimental and theoretical studies of the effect of the size, structure, and composition of NDs on the efficiency of reflectors. Methods of chemical and mechanical treatments of powders in order to modify their chemical composition and structure are discussed. The aim is to avoid, or at least to decrease, the neutron inelastic scatterers and absorbers (mainly hydrogen atoms but also metallic impurities and nitrogen) as well as to enhance coherent elastic scattering (to destroy ND clusters and sp2 carbon shells on the ND surface that result from the preparation of NDs). Issues requiring further study are identified. They include deeper purification of NDs from impurities that can be activated in high radiation fluxes, the stability of NDs in high radiation fluxes, and upscaling methods for producing larger quantities of ND powders. Possible ways of solving these problems are proposed. Full article
Show Figures

Figure 1

Figure 1
<p>The reflection probability for an isotropic neutron flux as a function of neutron velocity (bottom scale) and wavelength (top scale) for various carbon reflectors: (1) diamond-like coating (DLC) (thin solid line), (2) the best supermirror [<a href="#B62-nanomaterials-14-00387" class="html-bibr">62</a>] (dashed line), (3) hydrogen-free DND powder of infinite thickness (thick solid line), (4) calculation using the Monte Carlo N-Particle<sup>®</sup> (MCNP) standard program for reactor graphite reflector [<a href="#B44-nanomaterials-14-00387" class="html-bibr">44</a>] with infinite thickness at ambient temperature (dashed–dotted line). The reflection is elastic except for the last curve where only the fraction with no increase in energy is taken into account. Velocity (wavelength) ranges of ultracold, very cold, cold, and thermal neutrons are separated by broad gray vertical lines.</p>
Full article ">Figure 2
<p>Optimal diameters of diamond nanospheres (black line) as a function of VCN velocity (bottom scale) and wavelength (top scale). The solid red line shows the corresponding calculated VCN albedo for the optimal diameters of diamond nanospheres. Only neutron absorption by carbon is taken into account. The layer of DND is of the infinite thickness.</p>
Full article ">Figure 3
<p>A scheme (top view) of the installation used to measure neutron reflection/transmission from/through the ND layer. Two diaphragms (entry and exit) form the neutron beam in a horizontal plane. A mechanical neutron velocity selector is installed between the two diaphragms; it passes a band in the neutron velocity spectrum with a resolution of ∼<math display="inline"> <semantics> <mrow> <mn>20</mn> <mo>%</mo> </mrow> </semantics> </math>. The sample is placed in the neutron beam after the exit diaphragm so that the neutrons are incident along the normal to the plane of the sample, in the center of the detecting system, which consists of 11 neutron counters that detect the neutrons scattered on the sample. The geometry of the detecting system is such that the flat rectangular input windows of the counters are the side faces of a regular rectangular prism, at the base of which lies a regular dodecagon on a horizontal plane. The neutron beam passes through the center of one of the faces of the prism (free from the counter) perpendicular to this face. The input window of the counter, which forms the opposite face, measures the transmitted neutron beam. Each of the remaining ten detectors overlaps in the horizontal plane the corners with vertices in the center of the sample at ∼30° (∼28° taking into account the thickness of the side walls of the counter) and ∼60° in the vertical plane (the solid angle of ∼<math display="inline"> <semantics> <mrow> <mi>π</mi> <mo>/</mo> <mn>6</mn> </mrow> </semantics> </math>). The entire detecting system is surrounded by neutron shielding. To reduce scattering in the air, the entire detecting system (together with the sample placed at its center) was placed in an argon atmosphere.</p>
Full article ">Figure 4
<p>Size distribution of DNDs is provided by the producer (Ultradiamond, USA).</p>
Full article ">Figure 5
<p>The probability of detecting VCNs scattered on an Ultradiamond90 sample (normalized to the incident neutron flux) into the forward hemisphere (the sum of events in the detectors located at the angles of 30°, 60°, 300°, and 330°) and the rear hemisphere (the sum of events in the detectors located at the angles of 120°, 150°, 210°, and 240°) as a function of VCN velocity (bottom scale) and wavelength (top scale), for various powder thicknesses. Triangles correspond to the measured VCN backscattering probability. Circles stand for the VCN forward scattering probability. The dotted lines correspond to the model of independent nanoparticles at rest without taking into account the losses due to hydrogen impurities. Solid lines indicate the model taking into account the “heating” of VCNs on hydrogen impurities. Dotted lines correspond to the model ignoring the “heating” of VCNs on hydrogen impurities. The experimental scheme is shown in <a href="#nanomaterials-14-00387-f003" class="html-fig">Figure 3</a>. A similar figure is given in ref. [<a href="#B64-nanomaterials-14-00387" class="html-bibr">64</a>].</p>
Full article ">Figure 6
<p>A scheme of the VCN storage experiment [<a href="#B65-nanomaterials-14-00387" class="html-bibr">65</a>]. A collimated neutron beam of ∼1 cm in diameter enters the cylindrical cavity of the VCN trap with a diameter of ∼44 cm and a height of ∼47 cm, through a small hole (∼<math display="inline"> <semantics> <mrow> <mn>2</mn> <mo>×</mo> <mn>2</mn> </mrow> </semantics> </math><math display="inline"> <semantics> <mrow> <mspace width="3.33333pt"/> <mi>c</mi> <msup> <mi>m</mi> <mn>2</mn> </msup> </mrow> </semantics> </math>) in the side wall. VCNs, being repeatedly reflected from the VCN trap walls, can enter the exit hole with a diameter of ∼6 cm in the trap ceiling and be counted in a neutron detector located behind the hole. The incoming neutron beam can be closed or opened with a neutron valve. The velocity of VCNs in the beam is set with a velocity selector located in front of the neutron valve. The VCN trap is placed in a vacuum chamber with an input quartz window and an output thin aluminum window. If the VCN beam is closed, the count rate in the detector decreases exponentially, following the VCN density in the trap. Thus, one could measure the storage time of VCNs as a function of their velocity and wavelength. The experiment consists of measuring the decay time constant of the count rate in the detector after the neutron beam is closed at the entrance to the trap, as a function of the VCN velocity/wavelength. A similar figure is given in ref. [<a href="#B65-nanomaterials-14-00387" class="html-bibr">65</a>].</p>
Full article ">Figure 7
<p>A prototype of the side wall of the trap for storage of VCNs used in the experiment [<a href="#B65-nanomaterials-14-00387" class="html-bibr">65</a>]. A stage of manufacturing thin-walled tubes.</p>
Full article ">Figure 8
<p>Probability of VCN reflection from a layer of DNDs at ambient temperature as a function of their velocity and wavelength. Open circles correspond to the measurements at room temperature [<a href="#B65-nanomaterials-14-00387" class="html-bibr">65</a>]. Thin lines stand for the Monte Carlo calculations taking into account inelastic scattering of neutrons on hydrogen, with different values of inelastic scattering cross-section reduced to the neutron velocity of 2200 m/s: <math display="inline"> <semantics> <mrow> <mn>4</mn> <mspace width="3.33333pt"/> <mi>b</mi> </mrow> </semantics> </math> for pink dashed line, <math display="inline"> <semantics> <mrow> <mn>3</mn> <mspace width="3.33333pt"/> <mi>b</mi> </mrow> </semantics> </math> for red dashed line, <math display="inline"> <semantics> <mrow> <mn>1.3</mn><mspace width="3.33333pt"/> <mi>b</mi> </mrow> </semantics> </math> for dotted blue line, and <math display="inline"> <semantics> <mrow> <mn>0</mn> <mspace width="3.33333pt"/> <mi>b</mi> </mrow> </semantics> </math> for black line. A similar figure is given in ref. [<a href="#B65-nanomaterials-14-00387" class="html-bibr">65</a>].</p>
Full article ">Figure 9
<p>The reflection probability for an isotropic neutron flux as a function of neutron velocity (bottom scale) and wavelength (top scale) for various carbon reflectors: (1) diamond-like coating (DLC) (thin solid line), (2) the best supermirror [<a href="#B62-nanomaterials-14-00387" class="html-bibr">62</a>] (dashed line), (3) hydrogen-free DND powder of infinite thickness (thick solid line), (4) DND powder 3 cm thick at ambient temperature (points with error bars) with significant hydrogen contamination, (5) calculation using the Monte Carlo N-Particle<sup>®</sup> (MCNP) standard program for reactor graphite reflector [<a href="#B44-nanomaterials-14-00387" class="html-bibr">44</a>] with infinite thickness at ambient temperature (dashed-dotted line). The reflection is elastic except for the last curve where only the fraction with no increase in energy is taken into account. Velocity (wavelength) ranges of ultracold, very cold, cold, and thermal neutrons are separated by broad gray vertical lines.</p>
Full article ">Figure 10
<p>A scheme of the installation (top view) to measure QSR of neutrons. The neutron beam was shaped using two diaphragms. The angular beam divergence in the horizontal plane was ∼<math display="inline"> <semantics> <mrow> <mn>0.4</mn></mrow> </semantics> </math> mrad; the beam divergence in the vertical plane was ∼2°. The <math display="inline"> <semantics> <mi>λ</mi> </semantics> </math>-dependence measurements were carried out using the time-of-flight method (using a chopper). Reflected neutrons were recorded in a position-sensitive neutron detector. A similar figure is given in ref. [<a href="#B67-nanomaterials-14-00387" class="html-bibr">67</a>].</p>
Full article ">Figure 11
<p>The probability of neutron QSR from a DND sample within the solid angle of the detector as a function of the neutron wavelength (normalized to the incident neutron beam). The grazing angles <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> are 2°, 3°, and 4° (from top to bottom). Dark and open circles, as well as squares, correspond to the measured data; lines illustrate calculations (for the angles 2°, 3°, and 4° from top to bottom, respectively). A similar figure is given in ref. [<a href="#B66-nanomaterials-14-00387" class="html-bibr">66</a>].</p>
Full article ">Figure 12
<p>Measured angular distributions of QSR neutrons are shown with dark and open circles, and simulations are indicated by solid lines. The data are averaged over two ranges of incident neutron wavelengths: 0.4–0.5 nm and 0.8–0.9 nm, respectively. The curves are normalized to the incident neutron fluxes in the corresponding wavelength ranges. The neutron beam glancing angle is 2°. A similar figure is given in ref. [<a href="#B66-nanomaterials-14-00387" class="html-bibr">66</a>].</p>
Full article ">Figure 13
<p>A scheme for measuring the hydrogen content in powders. (1)—sample; (2)—neutron beam; (3)—collimator; (4)—germanium <math display="inline"> <semantics> <mi>γ</mi> </semantics> </math>-detector; (5)—<math display="inline"> <semantics> <mi>γ</mi> </semantics> </math> and neutron protection; (6)—monitor neutron detector. A similar figure is given in ref. [<a href="#B97-nanomaterials-14-00387" class="html-bibr">97</a>].</p>
Full article ">Figure 14
<p>The peak of total absorption in the reaction <math display="inline"> <semantics> <mrow> <mi>n</mi> <mo>(</mo> <mi>p</mi> <mi>d</mi> <mo>)</mo> <mi>γ</mi> </mrow> </semantics> </math> for polyethylene (squares), for Sample 1 (circuits) and Sample 2 (triangles) (see <a href="#nanomaterials-14-00387-t002" class="html-table">Table 2</a>). Spectra from an aluminum sample with a thickness equal to that of the aluminum containers are subtracted from the spectra measured with nanopowders. A similar figure is given in ref. [<a href="#B97-nanomaterials-14-00387" class="html-bibr">97</a>].</p>
Full article ">Figure 15
<p>(<b>a</b>) TEM image of an F-DND sample; (<b>b</b>) DND diameter distribution calculated from a set of similar images. The red dotted line corresponds to the lognormal distribution. The black solid line indicates the result of fitting the data with a smooth dependence. The figure is copied from ref. [<a href="#B73-nanomaterials-14-00387" class="html-bibr">73</a>].</p>
Full article ">Figure 16
<p>Examples of SEM images of DND (<b>left</b>) and F-DND (<b>right</b>). The figure is copied from ref. [<a href="#B73-nanomaterials-14-00387" class="html-bibr">73</a>].</p>
Full article ">Figure 17
<p>The probability density of the size distribution of scatterers as a function of their radius [nm] estimated using SANS (thick red dashed line), SAXS (thin dotted dash–two-dotted line), TEM (black thin dash–dotted line), and DLS (black thin solid line). The vertical black dotted line indicates the median radius as measured via radiography. All measurements were performed with F-DND samples, with the exception of SAXS, which was performed with a DND sample. A similar figure is given in ref. [<a href="#B73-nanomaterials-14-00387" class="html-bibr">73</a>].</p>
Full article ">Figure 18
<p>Comparison of the measured and simulated intensity <span class="html-italic">I</span> (cm<sup>−1</sup>) of scattered neutrons as a function of the transferred momentum <span class="html-italic">Q</span> (nm<sup>−1</sup>) for an F-DND sample. Black squares represent experimental data. The thin blue line shows the results of modeling within the model of discrete-size diamond nanoballs, and the thick red line additionally takes into account the background contribution at the level of <math display="inline"> <semantics> <mrow> <mn>9</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math> nm<sup>−1</sup>. A similar figure is given in ref. [<a href="#B73-nanomaterials-14-00387" class="html-bibr">73</a>].</p>
Full article ">Figure 19
<p>Probability density for the distribution function of nanoballs as a function of radius (in nm), estimated within the model of diamond nanoballs of a discrete set of sizes for the F-DND sample. Black circles show simulation results. The red solid line interpolates the simulation results. A similar figure is given in ref. [<a href="#B73-nanomaterials-14-00387" class="html-bibr">73</a>].</p>
Full article ">Figure 20
<p>Probability of neutron reflection (black lines) and absorption (red lines) for F-DND (solid lines) and DND (dash-dotted lines) depending on neutron velocity (on bottom) and wavelength (on top). The incident neutron flux is isotropic, the density of the powder is <math display="inline"> <semantics> <mrow> <mn>0.19</mn></mrow> </semantics> </math> g/cm<sup>3</sup>, and the thickness is infinite. A similar figure is given in ref. [<a href="#B20-nanomaterials-14-00387" class="html-bibr">20</a>].</p>
Full article ">Figure 21
<p>Size distributions of F-DNDs and DF-DNDs in ethanol, measured using the DLS method. The figure is copied from ref. [<a href="#B20-nanomaterials-14-00387" class="html-bibr">20</a>].</p>
Full article ">Figure 22
<p>Probability density as a function of radius (in nm) for the size distribution of nanoballs, calculated in the model of discrete-size diamond nanoballs for F-DND (solid circles) and DF-DND (open squares). The points correspond to the simulation. The figure is copied from ref. [<a href="#B20-nanomaterials-14-00387" class="html-bibr">20</a>].</p>
Full article ">Figure 23
<p>Neutron albedo for VCN velocities of 50 m/s (black lines), 100 m/s (red lines), and 150 m/s (blue lines) for F-DND (dashed lines) and DF-DND (solid lines) depending on the cavity radius. The incident neutron flux is isotropic, the powder thickness is infinite, and the densities of the DF-DND and F-DND are <math display="inline"> <semantics> <mrow> <mn>0.19</mn></mrow> </semantics> </math> g/cm<sup>3</sup> and <math display="inline"> <semantics> <mrow> <mn>0.56</mn></mrow> </semantics> </math> g/cm<sup>3</sup>, respectively. The figure is copied from ref. [<a href="#B20-nanomaterials-14-00387" class="html-bibr">20</a>].</p>
Full article ">Figure 24
<p>(<b>a</b>,<b>c</b>) Example TEM images of DF-DND and S-DND; (<b>b</b>,<b>d</b>) aize distributions for DF-DND and S-DND estimated using all available TEM images. The figure is copied from ref. [<a href="#B76-nanomaterials-14-00387" class="html-bibr">76</a>].</p>
Full article ">Figure 25
<p>Neutron scattering intensity <span class="html-italic">I</span> (cm<sup>−1</sup>) versus momentum transfer <span class="html-italic">Q</span> (nm<sup>−1</sup>) for DF-DND (black squares) and S-DND (green circles). To compare only the effect of particle size and not powder density, both curves are normalized to the sample density of 1 g/cm<sup>3</sup>. The figure is copied from ref. [<a href="#B76-nanomaterials-14-00387" class="html-bibr">76</a>].</p>
Full article ">Figure 26
<p>Neutron albedo for flat semi-infinite media DF-DND (dashed lines) and S-DND (solid lines) depending on the neutron velocity and wavelength. A similar figure is given in ref. [<a href="#B76-nanomaterials-14-00387" class="html-bibr">76</a>].</p>
Full article ">Figure 27
<p>Ratio of loss coefficients <math display="inline"> <semantics> <mrow> <msub> <mi>η</mi> <mrow> <mi>D</mi> <mi>F</mi> <mo>−</mo> <mi>D</mi> <mi>N</mi> <mi>D</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>η</mi> <mrow> <mi>S</mi> <mo>−</mo> <mi>D</mi> <mi>N</mi> <mi>D</mi> </mrow> </msub> </mrow> </semantics> </math> upon reflection from flat semi-infinite media of DF-DND and S-DND depending on the neutron velocity. The corresponding neutron albedo is shown in <a href="#nanomaterials-14-00387-f026" class="html-fig">Figure 26</a>. A similar figure is given in ref. [<a href="#B76-nanomaterials-14-00387" class="html-bibr">76</a>].</p>
Full article ">Figure 28
<p>The ratio of loss coefficients <math display="inline"> <semantics> <mrow> <msub> <mi>η</mi> <mrow> <mi>D</mi> <mi>F</mi> <mo>−</mo> <mi>D</mi> <mi>N</mi> <mi>D</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>η</mi> <mrow> <mi>S</mi> <mo>−</mo> <mi>D</mi> <mi>N</mi> <mi>D</mi> </mrow> </msub> </mrow> </semantics> </math> upon reflection from the wall of a spherical cavity with a radius of 5 cm and a wall thickness of 3 cm, for realistic powder densities of <math display="inline"> <semantics> <mrow> <mn>0.56</mn></mrow> </semantics> </math> g/cm<sup>3</sup> for DF-DND and <math display="inline"> <semantics> <mrow> <mn>0.67</mn></mrow> </semantics> </math> g/cm<sup>3</sup> for S-DND. A similar figure is given in ref. [<a href="#B76-nanomaterials-14-00387" class="html-bibr">76</a>].</p>
Full article ">
12 pages, 4463 KiB  
Article
The High Stability and Selectivity of Electrochemical Sensor Using Low-Cost Diamond Nanoparticles for the Detection of Anti-Cancer Drug Flutamide in Environmental Samples
by Nareshkumar Baskaran, Sanjay Ballur Prasanna, Yu-Chien Lin, Yeh-Fang Duann, Ren-Jei Chung and Yang Wei
Sensors 2024, 24(3), 985; https://doi.org/10.3390/s24030985 - 2 Feb 2024
Cited by 4 | Viewed by 1766
Abstract
In this study, a novel electrochemical sensor was created by fabricating a screen-printed carbon electrode with diamond nanoparticles (DNPs/SPCE). The successful development of the sensor enabled the specific detection of the anti-cancer drug flutamide (FLT). The DNPs/SPCE demonstrated excellent conductivity, remarkable electrocatalytic activity, [...] Read more.
In this study, a novel electrochemical sensor was created by fabricating a screen-printed carbon electrode with diamond nanoparticles (DNPs/SPCE). The successful development of the sensor enabled the specific detection of the anti-cancer drug flutamide (FLT). The DNPs/SPCE demonstrated excellent conductivity, remarkable electrocatalytic activity, and swift electron transfer, all of which contribute to the advantageous monitoring of FLT. These qualities are critical for monitoring FLT levels in environmental samples. Various structural and morphological characterization techniques were employed to validate the formation of the DNPs. Remarkably, the electrochemical sensor demonstrated a wide linear response range (0.025 to 606.65 μM). Additionally, it showed a low limit of detection (0.023 μM) and high sensitivity (0.403 μA μM−1 cm−2). Furthermore, the practicability of DNPs/SPCE can be successfully employed in FLT monitoring in water bodies (pond water and river water samples) with satisfactory recoveries. Full article
(This article belongs to the Special Issue Editorial Board Members' Collection Series: Aptamer Biosensors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns, (<b>b</b>) Raman spectra, (<b>c</b>) ball-stick, and (<b>d</b>) polyhedron structure of DNPs.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM image and (<b>b</b>) EDX analysis of DNPs.</p>
Full article ">Figure 3
<p>(<b>a</b>) EIS spectra of SPCE (Black dots) and DNPs-modified SPCE (Red dots) in 0.05 M [Fe (CN)<sub>6</sub>]<sup>3−/4−</sup> containing 0.1 M KCl (Inset: Randles equivalent circuit). (<b>b</b>) CV response of SPCE (with and without FLT), and DNPs/SPCE in 75 µM of FLT containing 0.1 M PBS (pH = 7.0). (<b>c</b>) CV responses with increasing FLT concentration (25 to 200 µM). (<b>d</b>) I<sub>pc</sub> vs. square root of the scan rate. (<b>e</b>) CV curves for scan rates (20 to 200 mVs<sup>−1</sup>) in 75 µM FLT (0.1 M, pH 7.0). (<b>f</b>) I<sub>pc</sub> vs. square root of scan rate and (<b>g</b>) Log I<sub>pc</sub> vs. log v. (<b>h</b>) CV curves of FLT (75 µM) at diverse pH (3, 5, 7, and 9). (<b>i</b>) values of I<sub>pc</sub> and E<sub>p</sub> versus pH.</p>
Full article ">Figure 4
<p>The possible electro-reduction of flutamide.</p>
Full article ">Figure 5
<p>(<b>a</b>) DPV responses for the various concentrations of FLT from 0.025 to 605.65 μM using the DNPs-modified SPCE (Inset: the calibration plot between the peak current vs. FLT concentration). (<b>b</b>) DPV curves for the DNPs/SPCE in the presence of FLT and FLT with various interfering substances. (<b>c</b>) Reproducibility and (<b>d</b>) storage stability.</p>
Full article ">Figure 6
<p>DPV response of (<b>a</b>) pond water and (<b>c</b>) river water samples containing FLT at different concentrations. (<b>b</b>,<b>d</b>) Corresponding linear plots.</p>
Full article ">
23 pages, 12377 KiB  
Article
Early Periods of Low-Temperature Linear Antenna CVD Nucleation and Growth Study of Nanocrystalline Diamond Films
by Awadesh Kumar Mallik, Wen-Ching Shih, Paulius Pobedinskas and Ken Haenen
Coatings 2024, 14(2), 184; https://doi.org/10.3390/coatings14020184 - 31 Jan 2024
Cited by 3 | Viewed by 1767
Abstract
Low-temperature growth of diamond films using the chemical vapor deposition (CVD) method is not so widely reported and its initial periods of nucleation and growth phenomenon are of particular interest to the researchers. Four sets of substrates were selected for growing diamond films [...] Read more.
Low-temperature growth of diamond films using the chemical vapor deposition (CVD) method is not so widely reported and its initial periods of nucleation and growth phenomenon are of particular interest to the researchers. Four sets of substrates were selected for growing diamond films using linear antenna microwave plasma-enhanced CVD (LA-MPCVD). Among them, silicon and sapphire substrates were pre-treated with detonation nanodiamond (DND) seeds before diamond growth, for enhancement of its nucleation. Carbon nanotube (CNT) films on Si substrates were also used as another template for LA-MPCVD diamond growth. To enhance diamond nucleation during CVD growth, some of the CNT films were again pre-treated by the electrophoretic deposition (EPD) of diamond nanoparticles. All these substrates were then put inside the LA-MPCVD chamber to grow diamond films under variable processing conditions. Microwave input powers (1100–2800 W), input power modes (pulse or continuous), antenna-to-stage distances (5–6.5 cm), process gas recipes (with or without CO2), methane gas percentages (3%–5%), and deposition times (11–120 min) were altered to investigate their effect on the growth of diamond film on the pre-treated substrates. The substrate temperatures were found to vary from as low as 170 °C to a maximum of 307 °C during the alteration of the different processing parameters. Contrary to the conventional MPCVD, it was observed that during the first hour of LA-MPCVD diamond growth, DND seeds and the nucleating structures do not coalesce together to make a continuous film. Deposition time was the most critical factor in fully covering the substrate surfaces with diamond film, since the substrate temperature could not become stable during the first hour of LA-MPCVD. CNTs were found to be oxidized rapidly under LA-MPCVD plasma conditions; therefore, a CO2-free process gas recipe was used to reduce CNT burning. Moreover, EPD-coated CNTs were found to be less oxidized by the LACVD plasma during diamond growth. Full article
(This article belongs to the Special Issue Chemical Vapor Deposition (CVD) of Coatings and Films)
Show Figures

Figure 1

Figure 1
<p>Four sets of substrates used in this study: (<b>a</b>) SEM image of the DND-seeded silicon substrate; (<b>b</b>) XRD of sapphire substrate (before DND seeding); (<b>c</b>) CNT-coated silicon substrate, inset SEM showing cross-sectional CNT length; and (<b>d</b>) top view of the EPD of diamond nanoparticles on the CNT/Si substrate.</p>
Full article ">Figure 2
<p>Computer screenshots during LA-MPCVD, showing (<b>a</b>) 60 min long plasma heating for sample # LA200519-1, (<b>b</b>) cooling–heating cycle for sample # LA200605-3, with 11 min of short CVD deposition, (<b>c</b>) stable profiles for 120 min long run with sample # LA200520-2, and (<b>d</b>) view of the microwave plasma heating of the Si, sapphire, CNT/Si, and EPD-diamond/CNT/Si substrates kept together.</p>
Full article ">Figure 3
<p>NCD crystals on DND-seeded silicon substrates. (<b>a</b>) SEM of the sample grown after first 15 min, (<b>b</b>) SEM of the sample grown after 30 min at 20k× magnification, (<b>c</b>) Raman spectra of the sample grown after 60 min, (<b>d</b>) SEM of the sample grown after 60 min, (<b>e</b>) SEM of the sample grown after 30 min at 200 k× magnification, and (<b>f</b>) Raman spectra of the sample grown after 15 min of LA-MPCVD growth.</p>
Full article ">Figure 4
<p>Raman signals from the NCD crystals grown over DND-seeded sapphire substrates after (<b>a</b>) 60, (<b>b</b>) 30, and (<b>c</b>) 15 min of LA-MPCVD growth at 1500 W power, 5 cm stage-to-antenna distance with gas recipe of H<sub>2</sub>/CH<sub>4</sub>/CO<sub>2</sub> = 89/5/6.</p>
Full article ">Figure 5
<p>Please refer to <a href="#coatings-14-00184-t001" class="html-table">Table 1</a>, row 4, process conditions, after 15 min of LA-MPCVD plasma treatment: (<b>a</b>) SEM at 500× magnification; (<b>b</b>) 20 k× magnification; and (<b>c</b>) Raman signal of EPD-diamond CNT/Si substrate after 15 min of LA-MPCVD plasma treatment. (<b>d</b>) SEM at 500× magnification, and (<b>e</b>) 20 k× magnification, and (<b>f</b>) Raman signal of CNT/Si substrate.</p>
Full article ">Figure 6
<p>Nearly coalesced NCD crystals grown using a LA-MPCVD reactor at a reduced power level of 1100 W on silicon substrate for 60 min: (<b>a</b>) 20 k× and (<b>b</b>) 50 k× magnification SEM images. Comparison of diamond nanocrystal coalescence on sapphire substrates after 60 min of LA-MPCVD at MW power levels of (<b>c</b>) 1500 W, isolated, and (<b>d</b>) 1100 W, connected.</p>
Full article ">Figure 7
<p>SEM images of the NCD films (sample # LA200525-1) grown using pulse mode MW (input 300 W/output 2000 W) power on Si substrates at (<b>a</b>) 5 k× and (<b>b</b>) 50 k× magnifications, and on sapphire substrates at (<b>c</b>) 5 kX and (<b>d</b>) 50 kX magnifications.</p>
Full article ">Figure 8
<p>Coalescence and secondary nucleation of NCDs. SEM of (<b>a</b>) interconnected NCDs on Si, pulse mode at 5 cm, (<b>b</b>) secondary nucleated NCDs on Si in pulse mode at 6.5 cm, (<b>c</b>) Raman spectra from secondary nucleated NCDs after 1 h LA-MPCVD in pulse mode at 6.5 cm. SEM images of NCDs after 2 h LA-MPCVD in CW mode at 6.5 cm, 1100 W power (<b>d</b>) on Si, and (<b>e</b>) on sapphire substrates, and (<b>f</b>) 50 k× magnification of image in 8e.</p>
Full article ">Figure 9
<p>Effect of LA-MPCVD plasma processing on EPD-seeded substrates, without CO<sub>2</sub> in the recipe. (<b>a</b>) Raman signal of the sample grown after 15 min, (<b>b</b>) Raman signal of the sample grown after 30 min, (<b>c</b>) top surface SEM at 20 k× magnification of the sample grown after 15 min, and (<b>d</b>) top surface SEM at 20 k× magnification of the sample grown after 30 min, with 500× inset image showing survival of CNTs after plasma treatment.</p>
Full article ">Figure 10
<p>LA-MPCVD growth of NCD crystals on DND-seeded silicon substrates without CO<sub>2</sub> in the recipe. SEM images showing early stages of CVD diamond growth at (<b>a</b>) 3%, (<b>b</b>) 4%, and (<b>c</b>) 5% CH<sub>4</sub> after 15 min of growth at an average input power of 1500 W. (<b>d</b>) Raman signal from such NCD crystals.</p>
Full article ">Figure 11
<p>Comparison of LA-MPCVD growth (1500 W, 5 cm, 15 min) of NCD crystals on DND-seeded silicon and sapphire substrates without CO<sub>2</sub> in the recipe. NCD crystal (sample # LA200605-1) SEM 50 k× magnification images on (<b>a</b>) Si and (<b>b</b>) sapphire substrates at 4% CH<sub>4</sub>. NCDs (sample # LA200605-2) on (<b>c</b>) Si and (<b>d</b>) sapphire substrates at 5% CH<sub>4</sub>.</p>
Full article ">Figure 12
<p>SEM images of LACVD-grown (1500 W, 5 cm) NCD crystals without CO<sub>2</sub> gas in the recipe, after 15 min on DND-seeded (<b>a</b>) silicon and (<b>b</b>) sapphire substrates and after 11 min on DND-seeded (<b>c</b>) silicon and (<b>d</b>) sapphire substrates.</p>
Full article ">Figure 13
<p>SEM images of LA-MPCVD-grown (1500 W) NCD crystals without CO<sub>2</sub> gas in the recipe, grown after 15 min on DND-seeded (<b>a</b>) silicon and (<b>b</b>) sapphire substrates at a 5 cm stage-to-antenna distance and grown after 30 min on DND-seeded (<b>c</b>) silicon and (<b>d</b>) sapphire substrates at a 6.5 cm stage-to-antenna distance.</p>
Full article ">
21 pages, 27250 KiB  
Article
Preparation and Properties of GO/ZnO/nHAp Composite Microsphere Bone Regeneration Material
by Jiang Wu, Chunmei Wang, Shuangsheng Zhang, Ling Zhang, Jingshun Hao, Zijian Jia, Xiaomei Zheng, Yuguang Lv, Shuang Fu and Guoliang Zhang
Micromachines 2024, 15(1), 122; https://doi.org/10.3390/mi15010122 - 11 Jan 2024
Cited by 3 | Viewed by 1548
Abstract
The purpose of this study is to explore the possibility of using graphene–zinc oxide–hydroxyapatite (GO/ZnO/nHAp) composite microspheres as bone regeneration materials by making use of the complementary advantages of nanocomposites, so as to provide reference for the clinical application of preventing and solving [...] Read more.
The purpose of this study is to explore the possibility of using graphene–zinc oxide–hydroxyapatite (GO/ZnO/nHAp) composite microspheres as bone regeneration materials by making use of the complementary advantages of nanocomposites, so as to provide reference for the clinical application of preventing and solving bacterial infection after implantation of synthetic materials. Firstly, GO/ZnO composites and hydroxyapatite nanoparticles were synthesized using the hydrothermal method, and then GO/ZnO/nHAp composite microspheres were prepared via high-temperature sintering. The graphene–zinc oxide–calcium phosphate composite microspheres were characterized by X-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM), X-ray photoelectron spectroscopy (XPS), energy dispersion spectroscopy (EDS), water contact angle measurement, degradation and pH determination, and differential thermal analysis (DiamondTG/DTA). The biocompatibility, osteogenic activity, and antibacterial activity of GO/ZnO/nHAp composite microspheres were further studied. The results of the cell experiment and antibacterial experiment showed that 0.5% and 1% GO-ZnO-nHAp composite microspheres not only had good biocompatibility and osteogenic ability but also inhibited Escherichia coli and Staphylococcus aureus by more than 45% and 70%. Therefore, GO/ZnO/nHAp composite microspheres have good physical and chemical properties and show good osteogenic induction and antibacterial activity, and this material has the possibility of being used as a bone regeneration material. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of synthesis and performance research of GO/ZnO/nHAp composite microspheres.</p>
Full article ">Figure 2
<p>HAp synthesis process.</p>
Full article ">Figure 3
<p>Macroscopic appearance of composite microspheres. (<b>a1</b>) Microspheres synthesized by ionic gel-drip method. (<b>a2</b>) Four groups of dried microspheres. (<b>a3</b>) After drying, the diameter of microspheres is about 1 mm.</p>
Full article ">Figure 4
<p>Growth state of MC3T3-E1 cells.</p>
Full article ">Figure 5
<p>Scanning electron microscopy of nHAp.</p>
Full article ">Figure 6
<p>XRD pattern of nHAp.</p>
Full article ">Figure 7
<p>FTIR pattern of nHAp.</p>
Full article ">Figure 8
<p>XRD pattern of composite microspheres.</p>
Full article ">Figure 9
<p>SEM images of composite microspheres: surface and sectional view. Surface diagram of each group of microspheres (<b>a1</b>,<b>b1</b>,<b>c1</b>,<b>d1</b>). Cross-sectional view of each group of microspheres (<b>a2</b>,<b>b2</b>,<b>c2</b>,<b>d2</b>).</p>
Full article ">Figure 10
<p>XPS spectra. (<b>a</b>,<b>b</b>) are the XPS spectra of the 1% group composite microspheres: (<b>a</b>) C 1 s; (<b>b</b>) Zn 2 p. (<b>c</b>,<b>d</b>) are the prepared GO-ZnO XPS spectra: (<b>c</b>) C 1 s; (<b>d</b>) Zn 2 p.</p>
Full article ">Figure 11
<p>EDS diagram of composite microspheres.</p>
Full article ">Figure 12
<p>Water contact angle test of composite microspheres.****, <span class="html-italic">p</span> &lt; 0.0001, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 13
<p>Changes in composite microspheres in PBS solution. (<b>a</b>) Degradation rate curve; (<b>b</b>) pH value change.</p>
Full article ">Figure 14
<p>TG curves of pure nHAp microspheres.</p>
Full article ">Figure 15
<p>The effect of different concentrations of composite microsphere extracts on the proliferation of MC3T3-E1 cells was detected by CCK-8, where (<b>a</b>) was 100% extract and (<b>b</b>) was 50% extract. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 16
<p>Effects of different concentrations of composite microsphere extracts on the activity of MC3T3-E1 cells ((<b>a</b>) is 100% extract, (<b>b</b>) is 50% extract); * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 17
<p>Effect of composite microsphere extract on the expression of ALP in MC3T3-E1 cells; * <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 18
<p>Colony count results: (<b>a</b>) <span class="html-italic">E. coli</span>, (<b>b</b>) <span class="html-italic">Staphylococcus aureus.</span></p>
Full article ">Figure 19
<p>Antibacterial rate of composite microspheres in each group: (<b>a</b>) <span class="html-italic">Staphylococcus aureus</span>; (<b>b</b>) <span class="html-italic">E. coli</span>; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 20
<p>Morphology of microspheres co-cultured with bacteria: (<b>a</b>) <span class="html-italic">E. coli</span>, (<b>b</b>) <span class="html-italic">Staphylococcus aureus.</span></p>
Full article ">Figure 21
<p>Bacterial growth curves: (<b>a</b>) <span class="html-italic">E. coli</span>, (<b>b</b>) <span class="html-italic">Staphylococcus aureus.</span></p>
Full article ">
11 pages, 5814 KiB  
Article
Multifunctional Core/Shell Diamond Nanoparticles Combining Unique Thermal and Light Properties for Future Biological Applications
by Sergey A. Grudinkin, Kirill V. Bogdanov, Vladimir A. Tolmachev, Mikhail A. Baranov, Ilya E. Kaliya, Valery G. Golubev and Alexander V. Baranov
Nanomaterials 2023, 13(24), 3124; https://doi.org/10.3390/nano13243124 - 12 Dec 2023
Viewed by 1350
Abstract
We report the development of multifunctional core/shell chemical vapor deposition diamond nanoparticles for the local photoinduced hyperthermia, thermometry, and fluorescent imaging. The diamond core heavily doped with boron is heated due to absorbed laser radiation and in turn heats the shell of a [...] Read more.
We report the development of multifunctional core/shell chemical vapor deposition diamond nanoparticles for the local photoinduced hyperthermia, thermometry, and fluorescent imaging. The diamond core heavily doped with boron is heated due to absorbed laser radiation and in turn heats the shell of a thin transparent diamond layer with embedded negatively charged SiV color centers emitting intense and narrowband zero-phonon lines with a temperature-dependent wavelength near 738 nm. The heating of the core/shell diamond nanoparticle is indicated by the temperature-induced spectral shift in the intensive zero-phonon line of the SiV color centers embedded in the diamond shell. The temperature of the core/shell diamond particles can be precisely manipulated by the power of the incident light. At laser power safe for biological systems, the photoinduced temperature of the core/shell diamond nanoparticles is high enough to be used for hyperthermia therapy and local nanothermometry, while the high zero-phonon line intensity of the SiV color centers allows for the fluorescent imaging of treated areas. Full article
(This article belongs to the Special Issue Biological Interactions of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Typical SEM images of the HFCVD diamond nanoparticles grown on SiO<sub>2</sub> opal surface: (<b>a</b>) a single diamond nanoparticle of ~1 µm in diameter; (<b>b</b>) an ensemble of the diamond nanoparticles with size of ~800 nm; (<b>c</b>) a single 1 µm diamond nanoparticle heavily doped with boron (BND particle).</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM image of the core/shell diamond particles of ca. 1.3 µm in diameter with a core heavily doped by boron (64,000 ppm in the gas mixture) and a shell with luminescent SiV centers on an opal surface. (<b>b</b>) A schematic image of hybrid core/shell diamond particle with a boron-doped core absorbing laser light and a diamond shell with luminescent SiV centers.</p>
Full article ">Figure 3
<p>Raman spectra of a BND particle doped with boron at a content in the gas mixture of 64,000 ppm on the opal substrate. Bands of heavily doped diamond (~1300 cm<sup>−1</sup>) and maximum in the region of ~1215 cm<sup>−1</sup>, corresponding to peaks in the single-phonon density of states of diamond, are marked.</p>
Full article ">Figure 4
<p>(<b>a</b>) Reflection spectra from diamond films: undoped (blue) and heavily doped with boron (red). The absence of interference fringes in the reflection spectra of the doped film is due to the significant absorption of radiation incident and reflected from the inner surface of the film due to the high boron content in the film. (<b>b</b>) Secondary emission of BND with a very weak luminescence background as compared to BND Raman peak at ~1213 cm<sup>−1</sup> (~518 nm). Excitation by light with by wavelength of ~488 nm.</p>
Full article ">Figure 5
<p>Raman and PL spectra of single core/shell ND with BND core and shell doped with SiV centers. Excitation wavelength is 488 nm. (<b>a</b>) Raman spectrum demonstrates several bands characteristic for slightly disordered lattice of diamond shell and a weak broadband of about 1220 cm<sup>−1</sup> from the BND core. The Raman shifts in the band are shown. (<b>b</b>) PL spectra of the core/shell diamond nanocrystal at two strongly different excitation powers of 0.08 and 0.8 mW, resulting in the spectral shift in the ZPL. Inset illustrates the ZPL shift from 738.9 nm to 740.1 nm. (<b>c</b>) A comparison between spectral positions of the SiV ZPLs for core/shell ND with absorbing BND core (red line) and ND without boron doping (blue line) obtained at the same excitation power of 0.8 mW.</p>
Full article ">Figure 6
<p>A set of PL spectra in the region of the ZPL of SiV centers excited by 532 nm radiation with different powers exerted on the sample. The vertical lines show maximum red shift in the ZPL peak with the increasing incident light power.</p>
Full article ">Figure 7
<p>(<b>a</b>) Dependence of the spectral position of SiV ZPL in CSNDs on the laser power; the inset shows the region of linear dependence. (<b>b</b>) The dependence of the core/shell ND temperature on the laser power, calculated using data on dependence of the SiV ZPL position on temperature from [<a href="#B20-nanomaterials-13-03124" class="html-bibr">20</a>]; the inset shows the region of linear dependence. (<b>c</b>) Dependence of SiV ZPL intensity of SiV center in CSNDs on laser radiation power, the region of linear dependence is shown. (<b>d</b>) Dependence of core/shell ND temperature on SiV ZPL position in actual power range for biological applications (black points). The obtained dependence is close to those obtained in the work of L. Golubewa et al. [<a href="#B17-nanomaterials-13-03124" class="html-bibr">17</a>] (blue points). Lines in insets are shown for convenience.</p>
Full article ">
28 pages, 3678 KiB  
Review
Nanoparticles and Mesenchymal Stem Cell (MSC) Therapy for Cancer Treatment: Focus on Nanocarriers and a si-RNA CXCR4 Chemokine Blocker as Strategies for Tumor Eradication In Vitro and In Vivo
by José Joaquín Merino and María Eugenia Cabaña-Muñoz
Micromachines 2023, 14(11), 2068; https://doi.org/10.3390/mi14112068 - 7 Nov 2023
Cited by 4 | Viewed by 3462
Abstract
Mesenchymal stem cells (MSCs) have a high tropism for the hypoxic microenvironment of tumors. The combination of nanoparticles in MSCs decreases tumor growth in vitro as well as in rodent models of cancers in vivo. Covalent conjugation of nanoparticles with the surface of [...] Read more.
Mesenchymal stem cells (MSCs) have a high tropism for the hypoxic microenvironment of tumors. The combination of nanoparticles in MSCs decreases tumor growth in vitro as well as in rodent models of cancers in vivo. Covalent conjugation of nanoparticles with the surface of MSCs can significantly increase the drug load delivery in tumor sites. Nanoparticle-based anti-angiogenic systems (gold, silica and silicates, diamond, silver, and copper) prevented tumor growth in vitro. For example, glycolic acid polyconjugates enhance nanoparticle drug delivery and have been reported in human MSCs. Labeling with fluorescent particles (coumarin-6 dye) identified tumor cells using fluorescence emission in tissues; the conjugation of different types of nanoparticles in MSCs ensured success and feasibility by tracking the migration and its intratumor detection using non-invasive imaging techniques. However, the biosafety and efficacy; long-term stability of nanoparticles, and the capacity for drug release must be improved for clinical implementation. In fact, MSCs are vehicles for drug delivery with nanoparticles and also show low toxicity but inefficient accumulation in tumor sites by clearance of reticuloendothelial organs. To solve these problems, the internalization or conjugation of drug-loaded nanoparticles should be improved in MSCs. Finally, CXCR4 may prove to be a promising target for immunotherapy and cancer treatment since the delivery of siRNA to knock down this alpha chemokine receptor or CXCR4 antagonism has been shown to disrupt tumor–stromal interactions. Full article
(This article belongs to the Special Issue Exploration and Application of Nanocarriers)
Show Figures

Figure 1

Figure 1
<p>Homing of mesenchymal stem cells (MSCs) toward the tumor site is regulated by CXCR4/SDF1 alpha levels. Inflammation is a calling signal for CXCR4 receptor-bearing MSCs and enhances the recruitment of MSCs [<a href="#B1-micromachines-14-02068" class="html-bibr">1</a>].</p>
Full article ">Figure 2
<p>Immunodulatory effects of mesenchymal SCs (MSCs).</p>
Full article ">Figure 3
<p>Mobilization of hematopoietic SCs (HSCs) by drugs (plerixaform, AMD-3100, CSF, or even viagre).</p>
Full article ">Figure 4
<p>Trophic factors released by MSCs contribute to cell adhesion, migration, survival, and senescence. For example, CXCR4 or CXCR5 chemokines increase MSC mobilization, while trophic factors such as HGF VGEF play a role in survival.</p>
Full article ">Figure 5
<p>Nanoparticle technology and physical methods for the introduction of several carriers in different types of cells (MSC, eritrocytes, etc…). Adapted from [<a href="#B87-micromachines-14-02068" class="html-bibr">87</a>].</p>
Full article ">Figure 6
<p>MSCs with antitumor drugs can eradicate a tumor using nanoparticle technology. Circulating MSCs can be targeted to the tumor by gene therapy when they overexpress the CXCR4 receptor by i.v injection in a rodent model of cancer, whereby tumor-reaching cells attempt to eradicate the tumor once released (Adapted from Pereboeva et al. [<a href="#B124-micromachines-14-02068" class="html-bibr">124</a>]).</p>
Full article ">
Back to TopTop