[go: up one dir, main page]

Next Issue
Volume 6, June
Previous Issue
Volume 6, April
 
 

Nanomaterials, Volume 6, Issue 5 (May 2016) – 14 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
4291 KiB  
Article
Effects of Atomization Injection on Nanoparticle Processing in Suspension Plasma Spray
by Hong-bing Xiong, Cheng-yu Zhang, Kai Zhang and Xue-ming Shao
Nanomaterials 2016, 6(5), 94; https://doi.org/10.3390/nano6050094 - 20 May 2016
Cited by 13 | Viewed by 8201
Abstract
Liquid atomization is applied in nanostructure dense coating technology to inject suspended nano-size powder materials into a suspension plasma spray (SPS) torch. This paper presents the effects of the atomization parameters on the nanoparticle processing. A numerical model was developed to simulate the [...] Read more.
Liquid atomization is applied in nanostructure dense coating technology to inject suspended nano-size powder materials into a suspension plasma spray (SPS) torch. This paper presents the effects of the atomization parameters on the nanoparticle processing. A numerical model was developed to simulate the dynamic behaviors of the suspension droplets, the solid nanoparticles or agglomerates, as well as the interactions between them and the plasma gas. The plasma gas was calculated as compressible, multi-component, turbulent jet flow in Eulerian scheme. The droplets and the solid particles were calculated as discrete Lagrangian entities, being tracked through the spray process. The motion and thermal histories of the particles were given in this paper and their release and melting status were observed. The key parameters of atomization, including droplet size, injection angle and velocity were also analyzed. The study revealed that the nanoparticle processing in SPS preferred small droplets with better atomization and less aggregation from suspension preparation. The injection angle and velocity influenced the nanoparticle release percentage. Small angle and low initial velocity might have more nanoparticles released. Besides, the melting percentage of nanoparticles and agglomerates were studied, and the critical droplet diameter to ensure solid melting was drawn. Results showed that most released nanoparticles were well melted, but the agglomerates might be totally melted, partially melted, or even not melted at all, mainly depending on the agglomerate size. For better coating quality, the suspension droplet size should be limited to a critical droplet diameter, which was inversely proportional to the cubic root of weight content, for given critical agglomerate diameter of being totally melted. Full article
(This article belongs to the Special Issue Plasma Nanoengineering and Nanofabrication)
Show Figures

Figure 1

Figure 1
<p>The schematic of liquid atomization.</p>
Full article ">Figure 2
<p>The schematic of suspension plasma spray with axially injected droplets or particles.</p>
Full article ">Figure 3
<p>Schematic of droplet secondary breakup.</p>
Full article ">Figure 4
<p>Schematic of droplet injection, solvent evaporation and nanoparticle release. Reproduced with permission from [<a href="#B17-nanomaterials-06-00094" class="html-bibr">17</a>]. Copyright Courtesy of Delbos, 2006.</p>
Full article ">Figure 5
<p>Comparison of gas temperature with published numerical (Reproduced with permission from [<a href="#B8-nanomaterials-06-00094" class="html-bibr">8</a>]. Copyright Courtesy of Jabbari, 2014) and experimental data (Reproduced with permission from [<a href="#B27-nanomaterials-06-00094" class="html-bibr">27</a>]. Copyright Courtesy of Brossa, 1988).</p>
Full article ">Figure 6
<p>Gas temperature contours in suspension plasma spray.</p>
Full article ">Figure 7
<p>Solvent evaporation position and nanoparticle release position for different droplet diameter.</p>
Full article ">Figure 8
<p>Solvent evaporation position and release position for different agglomerate diameter.</p>
Full article ">Figure 9
<p>Effects of injection angle on the nanoparticle release percentage.</p>
Full article ">Figure 10
<p>Effects of injection velocity on the nanoparticle release percentage.</p>
Full article ">Figure 11
<p>Velocity of nanoparticles and agglomerates.</p>
Full article ">Figure 12
<p>Temperature of nanoparticles and agglomerates.</p>
Full article ">
3075 KiB  
Article
Mechanochemical Synthesis of TiO2 Nanocomposites as Photocatalysts for Benzyl Alcohol Photo-Oxidation
by Weiyi Ouyang, Ewelina Kuna, Alfonso Yepez, Alina M. Balu, Antonio A. Romero, Juan Carlos Colmenares and Rafael Luque
Nanomaterials 2016, 6(5), 93; https://doi.org/10.3390/nano6050093 - 18 May 2016
Cited by 45 | Viewed by 8356
Abstract
TiO2 (anatase phase) has excellent photocatalytic performance and different methods have been reported to overcome its main limitation of high band gap energy. In this work, TiO2-magnetically-separable nanocomposites (MAGSNC) photocatalysts with different TiO2 loading were synthesized using a simple [...] Read more.
TiO2 (anatase phase) has excellent photocatalytic performance and different methods have been reported to overcome its main limitation of high band gap energy. In this work, TiO2-magnetically-separable nanocomposites (MAGSNC) photocatalysts with different TiO2 loading were synthesized using a simple one-pot mechanochemical method. Photocatalysts were characterized by a number of techniques and their photocatalytic activity was tested in the selective oxidation of benzyl alcohol to benzaldehyde. Extension of light absorption into the visible region was achieved upon titania incorporation. Results indicated that the photocatalytic activity increased with TiO2 loading on the catalysts, with moderate conversion (20%) at high benzaldehyde selectivity (84%) achieved for 5% TiO2-MAGSNC. These findings pointed out a potential strategy for the valorization of lignocellulosic-based biomass under visible light irradiation using designer photocatalytic nanomaterials. Full article
(This article belongs to the Special Issue Nanoparticles for Catalysis)
Show Figures

Figure 1

Figure 1
<p>Reaction system: (1) lamp cooling system; (2) double-walled immersion well reactor; (3) photoreactor; (4) port for taking samples; (5) 125 W ultraviolet (UV) lamp; (6) mother solution; and (7) magnetic stirrer.</p>
Full article ">Figure 2
<p>X-ray diffraction (XRD) pattern of 5% TiO<sub>2</sub>-magnetically-separable photocatalysts (MAGSNC) photocatalysts. (PDF 21-1272 and PDF 39-1346 are the card numbers for the crystalline structures in the data base, while Anatase, syn and Maghemite-C, syn are the corresponding structure names.)</p>
Full article ">Figure 3
<p>N<sub>2</sub> absorption-desorption isotherm of 5% TiO<sub>2</sub>-MAGSNC photocatalysts. P: partial vapor pressure of adsorbate gas in equilibrium with the surface at 77.4 K; P<sub>0</sub>: saturated pressure of adsorbate gas.</p>
Full article ">Figure 4
<p>Diffuse reflectance (DR) Ultraviolet-Visible (UV-VIS) absorption spectra of different TiO<sub>2</sub>-MAGSNC photocatalysts. P25: pure commercial TiO<sub>2</sub> from Evonik Industries.</p>
Full article ">Figure 5
<p>X-ray photoelectron spectroscopy (XPS) spectra of 5% TiO<sub>2</sub>-MAGSNC photocatalysts: (<b>a</b>) Ti 2p; and (<b>b</b>) Fe 2p.</p>
Full article ">Figure 6
<p>Scanning electron microscopy (SEM) images: (<b>a</b>) 5% TiO<sub>2</sub>-MAGSNC photocatalysts; (<b>b</b>) 2% TiO<sub>2</sub>-MAGSNC nanocomposites; and elements mapping of 2% TiO<sub>2</sub>-MAGSNC photocatalysts: (<b>c</b>) Si; (<b>d</b>) Fe; (<b>e</b>) Ti.</p>
Full article ">Figure 7
<p>Transmission electron microscopy (TEM) images of 5% TiO<sub>2</sub>-MAGSNC photocatalysts.</p>
Full article ">
3159 KiB  
Article
PDE5 Inhibitors-Loaded Nanovesicles: Physico-Chemical Properties and In Vitro Antiproliferative Activity
by Roberta F. De Rose, Maria Chiara Cristiano, Marilena Celano, Valentina Maggisano, Ada Vero, Giovanni Enrico Lombardo, Martina Di Francesco, Donatella Paolino, Diego Russo and Donato Cosco
Nanomaterials 2016, 6(5), 92; https://doi.org/10.3390/nano6050092 - 18 May 2016
Cited by 21 | Viewed by 5603
Abstract
Novel therapeutic approaches are required for the less differentiated thyroid cancers which are non-responsive to the current treatment. In this study we tested an innovative formulation of nanoliposomes containing sildenafil citrate or tadalafil, phosphodiesterase-5 inhibitors, on two human thyroid cancer cell lines (TPC-1 [...] Read more.
Novel therapeutic approaches are required for the less differentiated thyroid cancers which are non-responsive to the current treatment. In this study we tested an innovative formulation of nanoliposomes containing sildenafil citrate or tadalafil, phosphodiesterase-5 inhibitors, on two human thyroid cancer cell lines (TPC-1 and BCPAP). Nanoliposomes were prepared by the thin layer evaporation and extrusion methods, solubilizing the hydrophilic compound sildenafil citrate in the aqueous phase during the hydration step and dissolving the lipophilic tadalafil in the organic phase. Nanoliposomes, made up of 1,2-dipalmitoyl-sn-glycero-3-phosphatidylcholine monohydrate (DPPC), cholesterol, and N-(carbonyl-methoxypolyethylene glycol-2000)-1,2-distearoyl-sn-glycero-3-phosphoethanolamine (DSPE-mPEG2000) (6:3:1 molar ratio), were characterized by a mean diameter of ~100 nm, a very low polydispersity index (~0.1) and a negative surface charge. The drugs did not influence the physico-chemical properties of the systems and were efficiently retained in the colloidal structure. By using cell count and MTT assay, we found a significant reduction of the viability in both cell lines following 24 h treatment with both nanoliposomal-encapsulated drugs, notably greater than the effect of the free drugs. Our findings demonstrate that nanoliposomes increase the antiproliferative activity of phosphodiesterase-5 inhibitors, providing a useful novel formulation for the treatment of thyroid carcinoma. Full article
(This article belongs to the Special Issue Nanomaterials for Cancer Therapies)
Show Figures

Figure 1

Figure 1
<p>Transmission electron microscopy (TEM) micrographs of empty nanoliposomes (<b>A</b>); SIL-nanoliposomes (<b>B</b>) and TAD-nanoliposomes (<b>C</b>). Bar: 200 nm.</p>
Full article ">Figure 2
<p>Release profile of SIL and TAD from nanoliposomes. Experiments were carried out at room temperature. Values were the mean of three independent experiments in triplicate ± standard deviation (SD).</p>
Full article ">Figure 3
<p>Nanoliposomes containing PDE5 inhibitors reduced proliferation of TPC-1 cells. After 24 h of treatment, proliferation was evaluated by cell count assay (<b>A</b>) and MTT assay (<b>B</b>), as described in Materials and Methods. Results are mean ± SD of three experiments performed in triplicate. *, **, ***, <span class="html-italic">p</span> &lt; 0.05, 0.01, 0.001 <span class="html-italic">vs.</span> control (ctrl); °, °°, °°°, SIL-nlip <span class="html-italic">vs</span> SIL; <sup>#</sup>, <sup>##</sup>, <sup>###</sup>, <span class="html-italic">p</span> &lt; 0.05, 0.01, 0.001 TAD-nlip <span class="html-italic">vs.</span> TAD.</p>
Full article ">Figure 4
<p>Nanoliposomes containing PDE5 inhibitors reduced proliferation of BCPAP cells. After 24 h of treatment, proliferation was evaluated by cell count assay (<b>A</b>) and MTT assay (<b>B</b>), as described in Materials and Methods. Results are mean ± SD of three experiments performed in triplicate. *, **, ***, <span class="html-italic">p</span> &lt; 0.05, 0.01, 0.001 <span class="html-italic">vs.</span> control (ctrl); <sup>#</sup>, <sup>##</sup>, <span class="html-italic">p</span> &lt; 0.05, 0.01 TAD-nlip <span class="html-italic">vs.</span> TAD.</p>
Full article ">Figure 5
<p>Confocal laser scanning microscopy (CLSM) micrographs of TPC-1 cells treated with rhodamine-labelled nanoliposomes after 3 h incubation: <b>panel A</b>, Hoechst filter; <b>panel B</b>, TRITC filter; and <b>panel C</b>, overlay. A representative example of three independent experiments.</p>
Full article ">Figure 6
<p>Z-stack analysis by CLSM of TPC-1 cells treated with rhodamine-labelled nanoliposomes after 3 h incubation: <b>panel A</b>, Hoechst filter; <b>panel B</b>, TRITC filter; and <b>panel C</b>, overlay. A representative example of three independent experiments.</p>
Full article ">
5490 KiB  
Article
Post-Plasma SiOx Coatings of Metal and Metal Oxide Nanoparticles for Enhanced Thermal Stability and Tunable Photoactivity Applications
by Patrick Post, Nicolas Jidenko, Alfred P. Weber and Jean-Pascal Borra
Nanomaterials 2016, 6(5), 91; https://doi.org/10.3390/nano6050091 - 13 May 2016
Cited by 16 | Viewed by 6639
Abstract
The plasma-based aerosol process developed for the direct coating of particles in gases with silicon oxide in a continuous chemical vapor deposition (CVD) process is presented. It is shown that non-thermal plasma filaments induced in a dielectric barrier discharge (DBD) at atmospheric pressure [...] Read more.
The plasma-based aerosol process developed for the direct coating of particles in gases with silicon oxide in a continuous chemical vapor deposition (CVD) process is presented. It is shown that non-thermal plasma filaments induced in a dielectric barrier discharge (DBD) at atmospheric pressure trigger post-DBD gas phase reactions. DBD operating conditions are first scanned to produce ozone and dinitrogen pentoxide. In the selected conditions, these plasma species react with gaseous tetraethyl orthosilicate (TEOS) precursor downstream of the DBD. The gaseous intermediates then condense on the surface of nanoparticles and self-reactions lead to homogeneous solid SiOx coatings, with thickness from nanometer to micrometer. This confirms the interest of post-DBD injection of the organo-silicon precursor to achieve stable production of actives species with subsequent controlled thickness of SiOx coatings. SiOx coatings of spherical and agglomerated metal and metal oxide nanoparticles (Pt, CuO, TiO2) are achieved. In the selected DBD operating conditions, the thickness of homogeneous nanometer sized coatings of spherical nanoparticles depends on the reaction duration and on the precursor concentration. For agglomerates, operating conditions can be tuned to cover preferentially the interparticle contact zones between primary particles, shifting the sintering of platinum agglomerates to much higher temperatures than the usual sintering temperature. Potential applications for enhanced thermal stability and tunable photoactivity of coated agglomerates are presented. Full article
(This article belongs to the Special Issue Plasma Nanoengineering and Nanofabrication)
Show Figures

Figure 1

Figure 1
<p>Experimental setup with the different process steps.</p>
Full article ">Figure 2
<p>Effect of sintering on the morphology of platinum nanoparticles from the spark generator: (<b>left</b>) agglomerates downstream of the spark before sintering; and (<b>right</b>) spherical particles after sintering.</p>
Full article ">Figure 3
<p>Temporal evolution of voltage and dielectric barrier discharge (DBD) current (<b>left</b>) and DBD power <span class="html-italic">versus</span> peak-to-peak voltage characteristic of the DBD with insets of violet light emitted from plasmas in air (<b>right</b>).</p>
Full article ">Figure 4
<p>Mobility equivalent diameter measured with scanning mobility particles sizer (SMPS): (<b>left</b>) as a function of the tetraethyl orthosilicate (TEOS) concentration at constant reaction time <span class="html-italic">t</span><sub>Reaction</sub> = 83 s; and (<b>right</b>) as a function of reaction time at constant TEOS concentration of 0.7 ppmv (the total number concentration <span class="html-italic">N</span><sub>tot</sub> is given in the legend).</p>
Full article ">Figure 5
<p>Transmission electron microscopy (TEM) micrograph of final solid coating of nanoparticles for different applied voltage (<span class="html-italic">U</span><sub>pp</sub>): (<b>left</b>) 8 kV; (<b>middle</b>) 10 kV; and (<b>right</b>): 12 kV, for 0.4 ppmv TEOS and <span class="html-italic">t</span><sub>Reaction</sub> = 185 s.</p>
Full article ">Figure 6
<p>TEM micrographs: (<b>left</b>) Large SiO<sub><span class="html-italic">x</span></sub> particles (1.13 µm) formed in the gas phase during the reaction duration in the 20 L tank (<span class="html-italic">U</span><sub>pp</sub> = 10 kV, with post-DBD injection of <span class="html-italic">c</span><sub>TEOS</sub> = 1400 ppmv without Pt nanoparticles); (<b>middle</b>) smaller sub-micron sized SiO<sub><span class="html-italic">x</span></sub> coated Pt particles, formed and agglomerated in the gas phase (<span class="html-italic">U</span><sub>pp</sub> = 8.6 kV, 20 L tank with 280 ppmv TEOS and Pt nanoparticles); and (<b>right</b>) nanometer thick coatings of Pt nanoparticles (<span class="html-italic">U</span><sub>pp</sub> = 8 kV, <span class="html-italic">t</span><sub>Reaction</sub> = 185 s with 0.4 ppmv TEOS).</p>
Full article ">Figure 7
<p>TEM micrographs showing the time dependence of the coating thickness for a constant TEOS concentration of 0.7 ppmv.</p>
Full article ">Figure 8
<p>Coating thickness measured with scanning mobility particles sizer (SMPS) and TEM as a function of TEOS concentration for <span class="html-italic">t</span><sub>Reaction</sub> = 83 s (<b>left</b>) and as a function of the reaction time for 0.7 ppmv TEOS concentration (<b>right</b>).</p>
Full article ">Figure 9
<p>Influence of the particle material on the coating thickness from TEM analysis at a given TEOS concentration of 0.7 ppmv for <span class="html-italic">t</span><sub>Reaction</sub> = 83 s.</p>
Full article ">Figure 10
<p>Change in mobility equivalent diameter depending on the precursor concentration without (<b>left</b>) and with (<b>right</b>) plasma for agglomerates <span class="html-italic">t</span><sub>Reaction</sub> = 83 s).</p>
Full article ">Figure 11
<p>TEM micrograph of a coated Pt agglomerate after a reaction time of 185 s with a TEOS concentration of 0.7 ppmv showing a thick coating.</p>
Full article ">Figure 12
<p>Measurement of the coating coverage by photoemission for Pt particles and sketches of the different kinds of observed coatings from TEM micrographs (<span class="html-italic">t</span><sub>Reaction</sub> = 185 s). Additionally, the open, blue symbols show the change in APE signal depending on the sintering temperature, analogous to the particles shown in <a href="#nanomaterials-06-00091-f013" class="html-fig">Figure 13</a> (<span class="html-italic">t</span><sub>Reaction</sub> = 83 s).</p>
Full article ">Figure 13
<p>TEM micrographs showing the sintering behavior of non-coated (<b>top</b>) and coated (<b>bottom</b>) Pt particles during the transit through a tube furnace (residence time at 24 °C of 8 s).</p>
Full article ">
3459 KiB  
Article
Aggregation and Colloidal Stability of Commercially Available Al2O3 Nanoparticles in Aqueous Environments
by Julie Mui, Jennifer Ngo and Bojeong Kim
Nanomaterials 2016, 6(5), 90; https://doi.org/10.3390/nano6050090 - 13 May 2016
Cited by 52 | Viewed by 8259
Abstract
The aggregation and colloidal stability of three, commercially-available, gamma-aluminum oxide nanoparticles (γ-Al2O3 NPs) (nominally 5, 10, and 20–30 nm) were systematically examined as a function of pH, ionic strength, humic acid (HA) or clay minerals (e.g., montmorillonite) concentration using dynamic [...] Read more.
The aggregation and colloidal stability of three, commercially-available, gamma-aluminum oxide nanoparticles (γ-Al2O3 NPs) (nominally 5, 10, and 20–30 nm) were systematically examined as a function of pH, ionic strength, humic acid (HA) or clay minerals (e.g., montmorillonite) concentration using dynamic light scattering and transmission electron microscopy techniques. NPs possess pH-dependent surface charges, with a point of zero charge (PZC) of pH 7.5 to 8. When pH < PZC, γ-Al2O3 NPs are colloidally stable up to 100 mM NaCl and 30 mM CaCl2. However, significant aggregation of NPs is pronounced in both electrolytes at high ionic strength. In mixed systems, both HA and montmorillonite enhance NP colloidal stability through electrostatic interactions and steric hindrance when pH ≤ PZC, whereas their surface interactions are quite limited when pH > PZC. Even when pH approximates PZC, NPs became stable at a HA concentration of 1 mg·L−1. The magnitude of interactions and dominant sites of interaction (basal planes versus edge sites) are significantly dependent on pH because both NPs and montmorillonite have pH-dependent (conditional) surface charges. Thus, solution pH, ionic strength, and the presence of natural colloids greatly modify the surface conditions of commercial γ-Al2O3 NPs, affecting aggregation and colloidal stability significantly in the aqueous environment. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction (XRD) patterns of commercial γ-Al<sub>2</sub>O<sub>3</sub> nanoparticles (NPs) in size of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm. Note that 20–30 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs have a trace amount of boehmite (γ-AlOOH) (PDF 04-010-5683) marked with *, as an impurity.</p>
Full article ">Figure 2
<p>Bright field transmission electron microscopy (TEM) images of γ-Al<sub>2</sub>O<sub>3</sub> NPs in size of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm.</p>
Full article ">Figure 3
<p>Z-average (Z<sub>ave</sub>) hydrodynamic diameter (○) and ζ potential (●) of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of pH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Z<sub>ave</sub> hydrodynamic diameter of 5 nm (□), 10 nm (○), and 20–30 nm (△) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (NaCl); (<b>b</b>) Z<sub>ave</sub> hydrodynamic diameter of 5 nm (□), 10 nm (○), and 20–30 nm (△) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (CaCl<sub>2</sub>); (<b>c</b>) ζ potential of 5 nm (■), 10 nm (●), and 20–30 nm (▲) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (NaCl); and (<b>d</b>) ζ potential of 5 nm (■), 10 nm (●), and 20–30 nm (▲) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (CaCl<sub>2</sub>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Z<sub>ave</sub> hydrodynamic diameter (○) and ζ potential (●) of 10 nm of γ-Al<sub>2</sub>O<sub>3</sub> NPs as a function of HA concentration at pH close to the point of zero charge (PZC) of Al<sub>2</sub>O<sub>3</sub>; and (<b>b</b>) bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with 10 mg·L<sup>−1</sup> of HA at pH close to PZC of Al<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 6
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at pH &lt; PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">Figure 7
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">Figure 8
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at pH &gt; PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">
3760 KiB  
Article
Modification of the Interfacial Interaction between Carbon Fiber and Epoxy with Carbon Hybrid Materials
by Kejing Yu, Menglei Wang, Junqing Wu, Kun Qian, Jie Sun and Xuefeng Lu
Nanomaterials 2016, 6(5), 89; https://doi.org/10.3390/nano6050089 - 12 May 2016
Cited by 30 | Viewed by 5821
Abstract
The mechanical properties of the hybrid materials and epoxy and carbon fiber (CF) composites were improved significantly as compared to the CF composites made from unmodified epoxy. The reasons could be attributed to the strong interfacial interaction between the CF and the epoxy [...] Read more.
The mechanical properties of the hybrid materials and epoxy and carbon fiber (CF) composites were improved significantly as compared to the CF composites made from unmodified epoxy. The reasons could be attributed to the strong interfacial interaction between the CF and the epoxy composites for the existence of carbon nanomaterials. The microstructure and dispersion of carbon nanomaterials were characterized by transmission electron microscopy (TEM) and optical microscopy (OM). The results showed that the dispersion of the hybrid materials in the polymer was superior to other carbon nanomaterials. The high viscosity and shear stress characterized by a rheometer and the high interfacial friction and damping behavior characterized by dynamic mechanical analysis (DMA) indicated that the strong interfacial interaction was greatly improved between fibers and epoxy composites. Remarkably, the tensile tests presented that the CF composites with hybrid materials and epoxy composites have a better reinforcing and toughening effect on CF, which further verified the strong interfacial interaction between epoxy and CF for special structural hybrid materials. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The transmission electron microscopy (TEM) images of (<b>a</b>) hydroxyl graphene nanoplatelets (GnPs-OH); (<b>b</b>) hydroxyl multi-walled carbon nanotubes (MWCNTs-OH); (<b>c</b>) hybrid materials.</p>
Full article ">Figure 2
<p>Optical microscope images of tetrahydrofuran (THF) suspensions with (<b>a</b>) GnPs-OH; (<b>b</b>) MWCNTs-OH; (<b>c</b>) hybrid materials at a filler concentration of 0.1 mg/mL.</p>
Full article ">Figure 3
<p>The curing reaction curves of pure epoxy and epoxy composites with different fillers. EP: pure epoxy.</p>
Full article ">Figure 4
<p>The glass transition temperature (<span class="html-italic">T<sub>g</sub></span>) curves of pure epoxy and epoxy composites with different fillers.</p>
Full article ">Figure 5
<p>The viscosity <span class="html-italic">vs.</span> shear rate of epoxy and epoxy composites.</p>
Full article ">Figure 6
<p>The shear stress <span class="html-italic">vs.</span> shear rate of epoxy and epoxy composites.</p>
Full article ">Figure 7
<p>The dynamic mechanical thermal analysis (DMA) of (<b>A</b>) storage modulus (G′); (<b>B</b>) loss modulus (G″) and (<b>C</b>) dissipation factor (tanδ) <span class="html-italic">vs.</span> temperature.</p>
Full article ">Figure 8
<p>The tensile properties comparison of carbon fiber (CF) and epoxy composites (<b>a</b>) epoxy resin (EP)/CF; (<b>b</b>) GnPs-OH/EP/CF; (<b>c</b>) MWCNTs-OH/EP/CF; (<b>d</b>) hybrid materials/EP/CF.</p>
Full article ">Figure 9
<p>The tensile properties of CF composites.</p>
Full article ">Figure 10
<p>The SEM images of the fracture surface of CF composites (<b>a</b>) EP/CF; (<b>b</b>) MWCNTs-OH/EP/CF; (<b>c</b>) GnPs-OH/EP/CF; (<b>d</b>) hybrid materials/EP/CF.</p>
Full article ">
4198 KiB  
Article
Investigation of the Structural, Electrical, and Optical Properties of the Nano-Scale GZO Thin Films on Glass and Flexible Polyimide Substrates
by Fang-Hsing Wang, Kun-Neng Chen, Chao-Ming Hsu, Min-Chu Liu and Cheng-Fu Yang
Nanomaterials 2016, 6(5), 88; https://doi.org/10.3390/nano6050088 - 10 May 2016
Cited by 33 | Viewed by 7325
Abstract
In this study, Ga2O3-doped ZnO (GZO) thin films were deposited on glass and flexible polyimide (PI) substrates at room temperature (300 K), 373 K, and 473 K by the radio frequency (RF) magnetron sputtering method. After finding the deposition [...] Read more.
In this study, Ga2O3-doped ZnO (GZO) thin films were deposited on glass and flexible polyimide (PI) substrates at room temperature (300 K), 373 K, and 473 K by the radio frequency (RF) magnetron sputtering method. After finding the deposition rate, all the GZO thin films with a nano-scale thickness of about 150 ± 10 nm were controlled by the deposition time. X-ray diffraction patterns indicated that the GZO thin films were not amorphous and all exhibited the (002) peak, and field emission scanning electron microscopy showed that only nano-scale particles were observed. The dependences of the structural, electrical, and optical properties of the GZO thin films on different deposition temperatures and substrates were investigated. X-ray photoemission spectroscopy (XPS) was used to measure the elemental composition at the chemical and electronic states of the GZO thin films deposited on different substrates, which could be used to clarify the mechanism of difference in electrical properties of the GZO thin films. In this study, the XPS binding energy spectra of Ga2p3/2 and Ga2p1/2 peaks, Zn2p3/2 and Zn2p1/2 peaks, the Ga3d peak, and O1s peaks for GZO thin films on glass and PI substrates were well compared. Full article
Show Figures

Figure 1

Figure 1
<p>Surface morphologies of the Ga<sub>2</sub>O<sub>3</sub>-doped ZnO (GZO) thin films as a function of substrate and deposition temperature. (<b>a</b>) Glass; (<b>b</b>) polyimide (PI); (<b>c</b>) 373 K-Glass; (<b>d</b>) 373 K-PI; (<b>e</b>) 473 K-Glass; and (<b>f</b>) 473 K-PI, respectively.</p>
Full article ">Figure 2
<p>Atomic force microscopy (AFM) analysis of the GZO thin films as a function of substrate at room temperature. (<b>a</b>) Glass and (<b>b</b>) PI, respectively.</p>
Full article ">Figure 3
<p>X-ray diffraction (XRD) patterns of the undoped ZnO thin films and GZO thin films as a function of deposition temperature and on different substrates: (<b>a</b>) glass and (<b>b</b>) PI. a.u.: arbitrary unit.</p>
Full article ">Figure 4
<p>Transmittance spectra of the undoped ZnO thin films and GZO thin films deposited as a function of deposition temperature and different substrates: (<b>a</b>) glass and (<b>b</b>) PI.</p>
Full article ">Figure 5
<p>The (αhν)<sup>2</sup> <span class="html-italic">vs.</span> hν<span class="html-italic">-Eg</span> plots of the GZO thin films deposited as a function of deposition temperature and on different substrates: (<b>a</b>) glass and (<b>b</b>) PI.</p>
Full article ">Figure 6
<p>Variations of the (<b>a</b>) Hall mobility; (<b>b</b>) carrier concentration; and (<b>c</b>) resistivity of the GZO thin films as a function of used substrates and deposition temperature.</p>
Full article ">Figure 7
<p>Typical widescan spectrum of the GZO thin films as a function of used substrates. (<b>a</b>): glass; (<b>b</b>): PI.</p>
Full article ">Figure 8
<p>Binding energy spectra of (<b>a</b>) Ga2p<sub>3/2</sub> and Ga2p<sub>1/2</sub> peaks; (<b>b</b>) Zn2p<sub>3/2</sub> and Zn2p<sub>1/2</sub> peaks; and (<b>c</b>) Ga3d peak for the undoped ZnO thin films and GZO thin films on different substrates.</p>
Full article ">Figure 9
<p>Binding energy spectra of O<sub>1</sub><sub>s</sub> peaks for GZO thin films on different substrates: (<b>a</b>) glass and (<b>b</b>) PI.</p>
Full article ">
4376 KiB  
Article
Graphene FETs with Low-Resistance Hybrid Contacts for Improved High Frequency Performance
by Chowdhury Al-Amin, Mustafa Karabiyik, Phani Kiran Vabbina, Raju Sinha and Nezih Pala
Nanomaterials 2016, 6(5), 86; https://doi.org/10.3390/nano6050086 - 10 May 2016
Cited by 3 | Viewed by 5684
Abstract
This work proposes a novel geometry field effect transistor with graphene as a channel—graphene field-effect transistor (GFET), having a hybrid contact that consists of an ohmic source/drain and its extended part towards the gate, which is capacitively coupled to the channel. The ohmic [...] Read more.
This work proposes a novel geometry field effect transistor with graphene as a channel—graphene field-effect transistor (GFET), having a hybrid contact that consists of an ohmic source/drain and its extended part towards the gate, which is capacitively coupled to the channel. The ohmic contacts are used for direct current (DC) biasing, whereas their capacitive extension reduces access region length and provides the radio frequency (RF) signal a low impedance path. Minimization of the access region length, along with the paralleling of ohmic contact’s resistance and resistive part of capacitively coupled contact’s impedance, lower the overall source/drain resistance, which results in an increase in current gain cut-off frequency, fT. The DC and high-frequency characteristics of the two chosen conventional baseline GFETs, and their modified versions with proposed hybrid contacts, have been extensively studied, compared, and analyzed using numerical and analytical techniques. Full article
(This article belongs to the Special Issue 2D Nanomaterials: Graphene and Beyond Graphene)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Small signal equivalent circuit overlaid on top of a conventional graphene field-effect transistors (graphene FET).</p>
Full article ">Figure 2
<p>Small signal equivalent circuit overlaid on top of the proposed hybrid contact graphene FET.</p>
Full article ">Figure 3
<p>(<b>a</b>) Matching network-1; (<b>b</b>) matching network-2.</p>
Full article ">Figure 4
<p>Schematic of a radio frequency (RF) transmission line method (TLM) structure on graphene, with a small-signal equivalent circuit overlaid on top, and the equivalent two-port network.</p>
Full article ">Figure 5
<p>Schematic of baseline GFET-1 (not to scale).</p>
Full article ">Figure 6
<p>The I<sub>d</sub>-V<sub>d</sub> characteristics and I<sub>d</sub>-V<sub>g</sub> characteristics (inset) of the baseline GFET-1. RF characteristics (Current Gain, |h<sub>21</sub>| and Unilateral Power Gain, UPG) of the baseline GFET-1 plotted in decibel (dB) with respect to frequency.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic of the capacitor like structure; (<b>b</b>) The real and imaginary parts of impedance, estimated from simulations and analytical calculations.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic of RF TLM structure on graphene; (<b>b</b>) the real and imaginary part of C3 impedance estimated from both simulation and analytical calculations, plotted with respect to frequency.</p>
Full article ">Figure 9
<p>Schematic of the proposed GFET (not to scale).</p>
Full article ">Figure 10
<p>(<b>a</b>) Current Gain, |h21| of the baseline GFET-1 along with that of the proposed hybrid contact GFET in the electron regime (V<sub>gs</sub> = +2.0 V and V<sub>ds</sub> = +5.0 V) plotted with respect to frequency; (<b>b</b>) The current gain cut-off frequency (<span class="html-italic">f<sub>T</sub></span>) of the proposed GFET extracted from |h21| <span class="html-italic">vs.</span> <span class="html-italic">f</span> characteristics, plotted with respect to <span class="html-italic">L<sub>C3</sub></span>.</p>
Full article ">Figure 11
<p>(<b>a</b>) Current Gain, |h21| of the baseline GFET-2 along with that of the proposed hybrid contact GFET in the electron regime (V<sub>gs</sub> = +0.6 V and V<sub>ds</sub> = +1.6 V) plotted with respect to frequency; (<b>b</b>) The current gain cut-off frequency (<span class="html-italic">f<sub>T</sub></span>) of the proposed GFET extracted from |h21| <span class="html-italic">vs.</span> <span class="html-italic">f</span> characteristics, plotted with respect to <span class="html-italic">L<sub>C3</sub></span>.</p>
Full article ">
4050 KiB  
Article
Rapamycin Loaded Solid Lipid Nanoparticles as a New Tool to Deliver mTOR Inhibitors: Formulation and in Vitro Characterization
by Alice Polchi, Alessandro Magini, Jarosław Mazuryk, Brunella Tancini, Jacek Gapiński, Adam Patkowski, Stefano Giovagnoli and Carla Emiliani
Nanomaterials 2016, 6(5), 87; https://doi.org/10.3390/nano6050087 - 9 May 2016
Cited by 34 | Viewed by 7266
Abstract
Recently, the use of mammalian target of rapamycin (mTOR) inhibitors, in particular rapamycin (Rp), has been suggested to improve the treatment of neurodegenerative diseases. However, as Rp is a strong immunosuppressant, specific delivery to the brain has been postulated to avoid systemic exposure. [...] Read more.
Recently, the use of mammalian target of rapamycin (mTOR) inhibitors, in particular rapamycin (Rp), has been suggested to improve the treatment of neurodegenerative diseases. However, as Rp is a strong immunosuppressant, specific delivery to the brain has been postulated to avoid systemic exposure. In this work, we fabricated new Rp loaded solid lipid nanoparticles (Rp-SLN) stabilized with polysorbate 80 (PS80), comparing two different methods and lipids. The formulations were characterized by differential scanning calorimetry (DSC), nuclear magnetic resonance (NMR), wide angle X-ray scattering (WAXS), cryo-transmission electron microscopy (cryo-TEM), dynamic light scattering (DLS) and particle tracking. In vitro release and short-term stability were assessed. Biological behavior of Rp-SLN was tested in SH-SY5Y neuroblastoma cells. The inhibition of mTOR complex 1 (mTORC1) was evaluated over time by a pulse-chase study compared to free Rp and Rp nanocrystals. Compritol Rp-SLN resulted more stable and possessing proper size and surface properties with respect to cetyl palmitate Rp-SLN. Rapamycin was entrapped in an amorphous form in the solid lipid matrix that showed partial crystallinity with stable Lβ, sub-Lα and Lβ′ arrangements. PS80 was stably anchored on particle surface. No drug release was observed over 24 h and Rp-SLN had a higher cell uptake and a more sustained effect over a week. The mTORC1 inhibition was higher with Rp-SLN. Overall, compritol Rp-SLN show suitable characteristics and stability to be considered for further investigation as Rp brain delivery system. Full article
(This article belongs to the Special Issue Nanomaterials for Tissue Engineering)
Show Figures

Figure 1

Figure 1
<p>Hydrodynamic size and polydispersity change over time for Rp loaded compritol and cetyl palmitate SLN.</p>
Full article ">Figure 2
<p>Cryo-transmission electron microscopy (cryo-TEM) images of blank (<b>A</b>) and Rp-SLN (<b>B</b>). Magnification 180,000× and 200,000×.</p>
Full article ">Figure 3
<p>Particle tracking: dimensional profiles of blank and Rp-SLN at 25 °C (<b>A</b>) and an effect of temperature on the mean MHD of blank and Rp-SLN (<b>B</b>).</p>
Full article ">Figure 4
<p>Heating and cooling ramps of blank and Rp-SLN compared to bulk compritol and PS80.</p>
Full article ">Figure 5
<p>Proton nuclear magnetic resonance spectroscopy (<sup>1</sup>H NMR) spectra of (<b>A</b>) PS80; (<b>B</b>) blank and (<b>C</b>) Rp-SLN all prepared in D<sub>2</sub>O, were submitted to an external magnetic field of 18.8 T and <sup>1</sup>H resonance frequency of 800 MHz. The arrows indicate the area magnified in (<b>D</b>) corresponding to the PS80-derived oxyethylene moiety signals (about δ = 3.7) to highlight the slight shift occurring in SLN compared to pure PS80.</p>
Full article ">Figure 6
<p>WAXS profiles of pure Rp, compritol and blank and Rp-SLN. Arrows indicate the signals corresponding to the polymorphs observed for blank, Rp-SLN and bulk lipid.</p>
Full article ">Figure 7
<p>Cell uptake of SLN. Amount of Rp taken up by SH-SY5Y cells after 1, 2 and 4 h after the treatment with 200 nM Rp-SLN and Rp solution. * <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Cell uptake of fluorescent-SLN. Fluorescence microscopy images of SH-SY5Y cells were taken 1(<b>A</b>); 2 (<b>B</b>) and 4 h (<b>C</b>) after treatment with 500 nM DiQ-tagged Rp-SLN DiQ (red) and after staining of lysosomes with fluorescein isothiocyanate dextran (green). Nuclei were stained with 4',6-diamidino-2-phenylindole (DAPI). Magnification: 60×.</p>
Full article ">Figure 9
<p>Effect of the Rp treatments on cell proliferation. The Rp effect was evaluated in SH-SY5Y cells by pulse-chase experiments. Cells were seeded in a 96-well plate, incubated overnight at 37 °C and treated for 4 h with Rp solution (Rp-sol), Rp-nanocrystals (Rp-NC) and Rp-SLN at the concentration of 2, 10 and 20 nM (Pulse), panel (<b>A</b>–<b>C</b>), respectively. The cells were also treated with blank SLN as control. The cell proliferation was evaluated daily by using 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide (MTT) assay. Control (CTRL): untreated cells. Values are the mean ± S.D. of three independent experiments. * <span class="html-italic">p</span> &lt; 0.01 (Rp-SLN <span class="html-italic">vs.</span> blank SLN cells) according to unpaired two-tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 10
<p>Effect of Rp-SLN on mammalian target of rapamycin (mTOR) activity. The SH-SY5Y cells were seeded in a 6-well plate, incubated overnight at 37 °C and then treated for 4 h with 20 nM of Rp-sol or Rp-SLN (Pulse); untreated cells were considered as control (CTRL). (<b>A</b>) Cells were recovered after 0, 1, 2, 3 and 4 days (chase) and the immunoblotting analysis was performed for phospho-p70S6K (pThr389), p70S6K (Total) and actin. Representative Western blotting of three independent experiments is reported; (<b>B</b>) densitometric analysis of the immunoblot represents the percentage of the ratio between phospho-p70S6K (pThr389) with respect to p70S6K. Values are the mean ± S.D. of three independent experiments. * <span class="html-italic">p</span> &lt; 0.01 (Rp-SLN <span class="html-italic">vs.</span> CTRL cells) according to unpaired two tailed Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
2190 KiB  
Article
Rapid Nanoparticle Synthesis by Magnetic and Microwave Heating
by Viktor Chikan and Emily J. McLaurin
Nanomaterials 2016, 6(5), 85; https://doi.org/10.3390/nano6050085 - 5 May 2016
Cited by 71 | Viewed by 9227
Abstract
Traditional hot-injection (HI) syntheses of colloidal nanoparticles (NPs) allows good separation of the nucleation and growth stages of the reaction, a key limitation in obtaining monodisperse NPs, but with limited scalability. Here, two methods are presented for obtaining NPs via rapid heating: magnetic [...] Read more.
Traditional hot-injection (HI) syntheses of colloidal nanoparticles (NPs) allows good separation of the nucleation and growth stages of the reaction, a key limitation in obtaining monodisperse NPs, but with limited scalability. Here, two methods are presented for obtaining NPs via rapid heating: magnetic and microwave-assisted. Both of these techniques provide improved engineering control over the separation of nucleation and growth stages of nanomaterial synthesis when the reaction is initiated from room temperature. The advantages of these techniques with preliminary data are presented in this prospective article. It is shown here that microwave assisted heating could possibly provide some selectivity in activating the nanomaterial precursor materials, while magnetic heating can produce very tiny particles in a very short time (even on the millisecond timescale), which is important for scalability. The fast magnetic heating also allows for synthesizing larger particles with improved size distribution, therefore impacting, not only the quantity, but the quality of the nanomaterials. Full article
(This article belongs to the Special Issue Current Trends in Colloidal Nanocrystals)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) LaMer diagram for the nucleation an growth of nanocrystals. (<b>b</b>) Dependence of nucleation and growth rate of crystallization on monomer oversaturation S (S = 1 is the solubility of the monomer at any given temperature).</p>
Full article ">Figure 2
<p>Typical heating rates of magnetic, microwave and traditional heating methods. The heating rates depend on the power absorbed by the reaction mixture or contents.</p>
Full article ">Figure 3
<p>Drawing of two extremes of microwave-assisted reactions. The vial on the left contains a solution with a large tan δ (such as water). Upon microwave irradiation, the microwaves minimally penetrate the reaction volume due to efficient absorption by the solvent. On the right, a reaction with a solvent with a small tan δ (such as an alkane) is shown, and the microwaves can penetrate further into the reaction solution. The reactants in the solution have a larger tan δ, and can interact, increasing temperature.</p>
Full article ">Figure 4
<p>Photographs of a typical 10 mL microwave reaction vessel with a stir-bar and fiber-optic thermometer insert (<b>left</b>) and the fiber-optic thermometer (<b>right</b>). The glass insert provides values of the internal temperature, but is limited by response time and heat transfer through the insert.</p>
Full article ">Figure 5
<p>Temperature vs time plots of microwave heating of water (blue) and mineral spirits (orange) to 150 °C at 800 W. The dashed lines indicate the temperatures recorded by the internal fiber-optic thermometer and the solid lines are the temperatures read by the external infrared (IR) sensor.</p>
Full article ">Figure 6
<p>White light emission of ultrasmall quantum dots (QDs) produced in magnetic/microwave heating <span class="html-italic">vs.</span> other competing light sources (<b>a</b>). Absorption spectra of CdSe QDs synthesized when magnetic heating (<b>b</b>) is combined with traditional heating methods. The initial rapid magnetic heating provides a positive impact on the size distribution of the CdSe QDs.</p>
Full article ">
5810 KiB  
Review
Biosynthesis of Metal Nanoparticles: Novel Efficient Heterogeneous Nanocatalysts
by Jose M. Palomo and Marco Filice
Nanomaterials 2016, 6(5), 84; https://doi.org/10.3390/nano6050084 - 5 May 2016
Cited by 58 | Viewed by 7653
Abstract
This review compiles the most recent advances described in literature on the preparation of noble metal nanoparticles induced by biological entities. The use of different free or substituted carbohydrates, peptides, proteins, microorganisms or plants have been successfully applied as a new green concept [...] Read more.
This review compiles the most recent advances described in literature on the preparation of noble metal nanoparticles induced by biological entities. The use of different free or substituted carbohydrates, peptides, proteins, microorganisms or plants have been successfully applied as a new green concept in the development of innovative strategies to prepare these nanoparticles as different nanostructures with different forms and sizes. As a second part of this review, the application of their synthetic ability as new heterogonous catalysts has been described in C–C bond-forming reactions (as Suzuki, Heck, cycloaddition or multicomponent), oxidations and dynamic kinetic resolutions. Full article
(This article belongs to the Special Issue Nanoparticles for Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesis of gold nanoparticles (AuNPs) induced by different glucose derivatives.</p>
Full article ">Figure 2
<p>Characterization of the glucoside-induced AuNPs by high resolution transmission electron microscope (HRTEM).</p>
Full article ">Figure 3
<p>TEM image of different AgNPs. (<b>A</b>) AgNPs; (<b>B</b>) Mesoporous silica-coated silver nanoparticles (Ag@MSN) coated on a silicon substrate.</p>
Full article ">Figure 4
<p>Synthesis of glycosylated AuNPs. (<b>A</b>) Synthetic scheme; (<b>B</b>) TEM image of glyco-AgNPs.</p>
Full article ">Figure 5
<p>Self-organized AuNPs on the surfaces of biotin-Trp-Trp scaffold.</p>
Full article ">Figure 6
<p>Synthesis of AgNPs. (<b>A</b>) Peptide structure; (<b>B</b>,<b>C</b>) TEM pictures Ag<sup>+</sup>-peptide complex after different irradiation times, <span class="html-italic">t</span> = 30 s and <span class="html-italic">t</span> = 30 min, respectively.</p>
Full article ">Figure 7
<p>Synthesis of Biofunctionalized AuNPs. (<b>A</b>) Peptide and scheme of NPs formation; (<b>B</b>) TEM images of the AuNPs.</p>
Full article ">Figure 8
<p>Synthesis of Ag-biohybrid. (<b>A</b>) Scheme of the formation of the nanostructure; (<b>B</b>) TEM of Ag-nanohybrid. Reproduced with permission from [<a href="#B39-nanomaterials-06-00084" class="html-bibr">39</a>]. Copyright the Royal Society of Chemistry, 2013.</p>
Full article ">Figure 9
<p>Preparation of metal bionanohybrids.</p>
Full article ">Figure 10
<p>Eumelanine from tyrosinase induced the synthesis of gold nanoparticles. (<b>A</b>) TEM image of <span class="html-italic">R. etli</span> cell in the presence of 3-(3,4-dihydroxyphenyl)-<span class="html-small-caps">l</span>-alanine (<span class="html-small-caps">l</span>-DOPA) and Au ions; (<b>B</b>) TEM image of <span class="html-italic">R. etli</span> cell in the presence of Au ions; (<b>C</b>) TEM image of <span class="html-italic">R. etli</span> cell in the presence of <span class="html-small-caps">l</span>-DOPA; (<b>D</b>) TEM of the synthesized AuNPs. Reproduced with permission from [<a href="#B53-nanomaterials-06-00084" class="html-bibr">53</a>]. Copyright the Royal Society of Chemistry, 2014</p>
Full article ">Figure 11
<p>Synthesis of Pt nanoparticles induced by black wattle tannin (BWT). (<b>A</b>) Scheme of the formation of PtNPs by BWT; (<b>B</b>) TEM image of synthesized PtNPs using 2 mg of BWT; (<b>C</b>) TEM image of synthesized PtNPs using 15 mg of BWT. Reproduced with permission from [<a href="#B54-nanomaterials-06-00084" class="html-bibr">54</a>]. Copyright the Royal Society of Chemistry, 2016.</p>
Full article ">Figure 12
<p>Characterization of Fe-Pd NPs synthesized by <span class="html-italic">Ulmus</span> <span class="html-italic">davidiana.</span> (<b>A</b>) TEM image of Fe-Pd NPs; (<b>B</b>) Particle size distribution histogram. Reproduced with permission from [<a href="#B12-nanomaterials-06-00084" class="html-bibr">12</a>]. Copyright the Royal Society of Chemistry, 2015.</p>
Full article ">Figure 13
<p>Synthesis of substituted carboxamides catalyzed by Fe-Pd NPs nanocatalyst.</p>
Full article ">Figure 14
<p>Oxidation of phenylalcohols by BWT-PtNPs.</p>
Full article ">Figure 15
<p>Multicomponent reaction catalyzed by AuNPs Hybrid.</p>
Full article ">Figure 16
<p>Catalytic applications of CAL-B-PdNPs hybrid.</p>
Full article ">
5217 KiB  
Article
Noise Removal with Maintained Spatial Resolution in Raman Images of Cells Exposed to Submicron Polystyrene Particles
by Linnea Ahlinder, Susanne Wiklund Lindström, Christian Lejon, Paul Geladi and Lars Österlund
Nanomaterials 2016, 6(5), 83; https://doi.org/10.3390/nano6050083 - 29 Apr 2016
Cited by 8 | Viewed by 5283
Abstract
The biodistribution of 300 nm polystyrene particles in A549 lung epithelial cells has been studied with confocal Raman spectroscopy. This is a label-free method in which particles and cells can be imaged without using dyes or fluorescent labels. The main drawback with Raman [...] Read more.
The biodistribution of 300 nm polystyrene particles in A549 lung epithelial cells has been studied with confocal Raman spectroscopy. This is a label-free method in which particles and cells can be imaged without using dyes or fluorescent labels. The main drawback with Raman imaging is the comparatively low spatial resolution, which is aggravated in heterogeneous systems such as biological samples, which in addition often require long measurement times because of their weak Raman signal. Long measurement times may however induce laser-induced damage. In this study we use a super-resolution algorithm with Tikhonov regularization, intended to improve the image quality without demanding an increased number of collected pixels. Images of cells exposed to polystyrene particles have been acquired with two different step lengths, i.e., the distance between pixels, and compared to each other and to corresponding images treated with the super-resolution algorithm. It is shown that the resolution after application of super-resolution algorithms is not significantly improved compared to the theoretical limit for optical microscopy. However, to reduce noise and artefacts in the hyperspectral Raman images while maintaining the spatial resolution, we show that it is advantageous to use short mapping step lengths and super-resolution algorithms with appropriate regularization. The proposed methodology should be generally applicable for Raman imaging of biological samples and other photo-sensitive samples. Full article
(This article belongs to the Special Issue Nanoparticles in Bioimaging)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Loadings for principal components (PCs) ((<b>a</b>) PC1 and (<b>b</b>) PC2) for principal component analysis (PCA) of the 500 nm image. Loadings for (<b>c</b>) PC1 and (<b>d</b>) PC2 for PCA of the 100 nm image. Raman bands assigned to polystyrene submicron (PS) particles are marked with dashed lines.</p>
Full article ">Figure 2
<p>Loadings for PC2 for (<b>a</b>) the 500 nm image; and (<b>b</b>) the 100 nm image. (<b>c</b>) Reference spectrum of PS multiplied by −1. Raman bands assigned to PS are marked with dashed lines. An intensity offset have been added to the spectra.</p>
Full article ">Figure 3
<p>Normalized score maps of PC1 from the PCA model of (<b>a</b>) the 100 nm image and (<b>b</b>) the 500 nm image. Normalized score maps of PC2 from the PCA of (<b>c</b>) the 100 nm image and (<b>d</b>) the 500 nm image. The large crosses (lines at <span class="html-italic">X</span> ≈ 11.5 µm and <span class="html-italic">Y</span> ≈ 16.1 µm in the 100 nm image and at <span class="html-italic">X</span> ≈ 8.5 µm and <span class="html-italic">Y</span> ≈ 21 µm in the 500 nm image) and small crosses show the geometrical centers of large and small particle agglomerates, respectively. Score values in image b) have been multiplied with −1 for easier comparison. The maps are slightly offset by ≈ 3 µm (<span class="html-italic">X</span>) and ≈ 5 µm (<span class="html-italic">Y</span>) and the (0,0) position is thus not the same in the 100 nm image and the 500 nm image. The 100 nm image covers a mapping area of 24.4 × 32.5 µm<sup>2</sup>, and the 500 nm image 28.5 × 32.5 µm<sup>2</sup>.</p>
Full article ">Figure 4
<p>Raw spectra from (<b>a</b>) a pixel centered at the largest particle agglomerate; (<b>b</b>) a pixel from the cell in a region free from particles; and (<b>c</b>) the background (substrate). All pixels are from the 100 nm image. An intensity offset have been added to the spectra.</p>
Full article ">Figure 5
<p>Optical microscope image of the Raman mapped cell. The cell has been exposed to PS particles.</p>
Full article ">Figure 6
<p>Optical microscope image of a cell that has not been exposed to PS. Black dots are normally visible inside cells although they have not been exposed to PS.</p>
Full article ">Figure 7
<p>Point spread function (PSF) approximated by fitting Gaussian functions to the derivatives of line maps from measurements across the edge of an Au structure in a 1951 USAF patterned Au/Si reference sample.</p>
Full article ">Figure 8
<p>Intensity line scans of score-values from thePC2 score maps in <a href="#nanomaterials-06-00083-f003" class="html-fig">Figure 3</a>c,d after application of the super-resolution algorithm with different α-values at different location in the images where particles are located: (<b>a</b>) <span class="html-italic">X</span> = 8.5 µm in PC2 for the 500 nm image; (<b>b</b>) <span class="html-italic">Y</span> = 21 µm in the 500 nm image; (<b>c</b>) <span class="html-italic">X</span> = 11.5 µm, in the 100 nm image; and (<b>d</b>) <span class="html-italic">Y</span> = 16.1 µm in the 100 nm image. The curves are shifted along the ordinate axis for clarity.</p>
Full article ">Figure 9
<p>Normalized images of PC2, <span class="html-italic">i.e.</span>, score values ranging from 0 to 1. (<b>a</b>) 500 nm image; (<b>b</b>) 500 nm image treated with super-resolution algorithm, α = 0.05; (<b>c</b>) 500 nm image treated with super-resolution, α = 0.5; (<b>d</b>) 100 nm image; (<b>e</b>) 100 nm image treated with super-resolution algorithm α = 0.05; (<b>f</b>) 100 nm image treated with super-resolution algorithm, α = 0.5. Note that the (0,0) point is not the same in the 500 nm images and the 100 nm images.</p>
Full article ">Figure 10
<p>Intensity profiles for lines at X = 16.5 µm from the PC2 score (t[2]) map from the (<b>a</b>) 500 nm image; (<b>b</b>) super-resolution 500 nm image, α = 0.05; and (<b>c</b>) super-resolution 500 nm image, α = 0.5 Intensity profiles at <span class="html-italic">X</span> = 11.5 µm from the PC2 score (t[2]) map of the (<b>d</b>) 100 nm image; (<b>e</b>) super-resolution 100 nm image, α = 0.05; and (<b>f</b>) super-resolution 100 nm image, α = 0.5.</p>
Full article ">Figure 11
<p>Threshold images of the score map for PC2 of (<b>a</b>) 500 nm image; (<b>b</b>) super-resolution 500 nm image, α = 0.05; (<b>c</b>) super-resolution 500 nm image, α = 0.5; (<b>d</b>) 100 nm image; (<b>e</b>) super-resolution 100 nm image, α = 0.05; (<b>f</b>) super-resolution 100 nm image, α = 0.5. The arrows point at a particle agglomerate that is difficult to detect without applying super-resolution. Threshold values are calculated with the “triangle method” [<a href="#B24-nanomaterials-06-00083" class="html-bibr">24</a>].</p>
Full article ">
3888 KiB  
Article
Application of L-Aspartic Acid-Capped ZnS:Mn Colloidal Nanocrystals as a Photosensor for the Detection of Copper (II) Ions in Aqueous Solution
by Jungho Heo and Cheong-Soo Hwang
Nanomaterials 2016, 6(5), 82; https://doi.org/10.3390/nano6050082 - 27 Apr 2016
Cited by 10 | Viewed by 5790
Abstract
Water-dispersible ZnS:Mn nanocrystals (NCs) were synthesized by capping the surface with polar L-aspartic acid (Asp) molecules. The obtained ZnS:Mn-Asp NC product was optically and physically characterized using the corresponding spectroscopic methods. The ultra violet-visible (UV-VIS) absorption spectrum and photoluminescence (PL) emission spectrum of [...] Read more.
Water-dispersible ZnS:Mn nanocrystals (NCs) were synthesized by capping the surface with polar L-aspartic acid (Asp) molecules. The obtained ZnS:Mn-Asp NC product was optically and physically characterized using the corresponding spectroscopic methods. The ultra violet-visible (UV-VIS) absorption spectrum and photoluminescence (PL) emission spectrum of the NCs showed broad peaks at 320 and 590 nm, respectively. The average particle size measured from the obtained high resolution-transmission electron microscopy (HR-TEM) image was 5.25 nm, which was also in accordance with the Debye-Scherrer calculations using the X-ray diffraction (XRD) data. Moreover, the surface charge and degree of aggregation of the ZnS:Mn-Asp NCs were determined by electrophoretic and hydrodynamic light scattering methods, respectively. These results indicated the formation of agglomerates in water with an average size of 19.8 nm, and a negative surface charge (−4.58 mV) in water at ambient temperature. The negatively-charged NCs were applied as a photosensor for the detection of specific cations in aqueous solution. Accordingly, the ZnS:Mn-Asp NCs showed an exclusive luminescence quenching upon addition of copper (II) cations. The kinetic mechanism study on the luminescence quenching of the NCs by the addition of the Cu2+ ions proposed an energy transfer through the ionic binding between the two oppositely-charged ZnS:Mn-Asp NCs and Cu2+ ions. Full article
(This article belongs to the Special Issue Current Trends in Colloidal Nanocrystals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>High resolution-transmission electron microscopy (HR-TEM) images of the ZnS:Mn-Asp nanocrystals (NCs). Asp: L-aspartic acid.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ultraviolet-visible (UV-VIS) absorption spectrum; and (<b>b</b>) room temperature solution photoluminescence (PL) emission spectrum of the ZnS:Mn-Asp NCs. (λ<sub>max</sub> of the fluorescence peak was 590 nm).</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Ultraviolet-visible (UV-VIS) absorption spectrum; and (<b>b</b>) room temperature solution photoluminescence (PL) emission spectrum of the ZnS:Mn-Asp NCs. (λ<sub>max</sub> of the fluorescence peak was 590 nm).</p>
Full article ">Figure 3
<p>X-ray diffraction (XRD) patterns of: (<b>a</b>) ZnS:Mn-Asp NCs; and (<b>b</b>) ZnS bulk solid in a cubic zinc-blende phase (JCPDS 05-0566) as an internal reference.</p>
Full article ">Figure 4
<p>Fourier transform infrared (FT-IR) spectrum of the ZnS:Mn-Asp NCs.</p>
Full article ">Figure 5
<p>Light scattering image of the ZnS:Mn-Asp NCs upon addition of transition metal ions, taken under irradiation of He-Cd laser (325 nm) light. (‘Blank’ refers to ZnS:Mn-Asp NCs only).</p>
Full article ">Figure 6
<p>Emission (PL) spectra of the ZnS:Mn-Asp NCs following the addition of divalent transition metal ions. The emission peak of the ZnS:Mn-Asp-Cu (brown) was exclusively diminished. (‘Blank’ refers to no addition of transition metal ions).</p>
Full article ">Figure 7
<p>A linear fitting diagram showing the modified (natural log-dependent) Stern-Volmer relationship between the ZnS:Mn-Asp NCs and the added Cu<sup>2+</sup> ions. (<span class="html-italic">R</span><sup>2</sup> = 0.9989, and the quenching rate constant k = 3.14 × 10<sup>4</sup> M<sup>−1</sup>).</p>
Full article ">Figure 8
<p>Particle size distribution histograms of the ZnS:Mn-Asp NCs only (black, 19.8 nm in average), and the ZnS:Mn-Asp-Cu adduct (green, 58.4 nm in average) in aqueous solution.</p>
Full article ">Figure 9
<p>X-band electron paramagnetic resonance (EPR) spectrum of the ZnS:Mn-Asp-Cu adduct (solid) taken at 200 K. The spectrum of a paramagnetic CuCl<sub>2</sub> (d<sup>9</sup>) solid (dashes) is also presented for comparison.</p>
Full article ">
3701 KiB  
Review
Dye-Doped Fluorescent Silica Nanoparticles for Live Cell and In Vivo Bioimaging
by Wen-Han Zhang, Xiao-Xiao Hu and Xiao-Bing Zhang
Nanomaterials 2016, 6(5), 81; https://doi.org/10.3390/nano6050081 - 27 Apr 2016
Cited by 67 | Viewed by 11135
Abstract
The need for novel design strategies for fluorescent nanomaterials to improve our understanding of biological activities at the molecular level is increasing rapidly. Dye-doped fluorescent silica nanoparticles (SiNPs) emerge with great potential for developing fluorescence imaging techniques as a novel and ideal platform [...] Read more.
The need for novel design strategies for fluorescent nanomaterials to improve our understanding of biological activities at the molecular level is increasing rapidly. Dye-doped fluorescent silica nanoparticles (SiNPs) emerge with great potential for developing fluorescence imaging techniques as a novel and ideal platform for the monitoring of living cells and the whole body. Organic dye-containing fluorescent SiNPs exhibit many advantages: they have excellent biocompatibility, are non-toxic, highly hydrophilic, optically transparent, size-tunable and easily modified with various biomolecules. The outer silica shell matrix protects fluorophores from outside chemical reaction factors and provides a hydrophilic shell for the insoluble nanoparticles, which enhances the photo-stability and biocompatibility of the organic fluorescent dyes. Here, we give a summary of the synthesis, characteristics and applications of fluorescent SiNPs for non-invasive fluorescence bioimaging in live cells and in vivo. Additionally, the challenges and perspectives of SiNPs are also discussed. We prospect that the further development of these nanoparticles will lead to an exciting breakthrough in the understanding of biological processes. Full article
(This article belongs to the Special Issue Nanoparticles in Bioimaging)
Show Figures

Figure 1

Figure 1
<p>Simple procedure synthesis of mesoporous silica nanoparticles.</p>
Full article ">Figure 2
<p>(<b>A</b>) Procedures for preparation amine group- and TAT peptide-conjugated mesoporous silica nanoparticles (MSNs); (<b>B</b>) an illustration of doxorubicin (Dox)@MSNs-TAT for targeting nuclei of cancer cells and delivering/releasing drugs directly into nuclei; (<b>C</b>) Transmission electron microscope transmission electron microscope (TEM) images of MSNs with different sizes. Scale bars: 100 nm. Reproduced with permission from [<a href="#B55-nanomaterials-06-00081" class="html-bibr">55</a>]. Copyright 2012 American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>A</b>) Chemical reaction routes of pH-responsive hierarchical poreSiNP (HPSN)-<span class="html-italic">N</span>,<span class="html-italic">N</span>-phenylenebis(salicylideneimine)dicarboxylic acid (Salphdc)-folate (FA) nanosystem; (<b>B</b>) the tumor therapy and bioimaging <span class="html-italic">in vivo</span> of the drug-loaded HPSN-Salphdc-FA nanosystem. Reproduced with permission from [<a href="#B39-nanomaterials-06-00081" class="html-bibr">39</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>A</b>) Whole-body real-time fluorescence imaging of DOX@HPSN-Salphdc (I) and DOX@HPSN-Salphdc-FA (II) at different hours. Scale bars: 3 cm. (<b>B</b>) Histogram of the fluorescence intensity of tumors issue treated with Dox@HPSN-Salphdc and Dox@HPSN-Salphdc-FA at each interval, respectively. (<b>C</b>) Images of mainly organs and (<b>D</b>) quantitative energy dispersive spectrometry analysis after injection of DOX@HPSN-Salphdc-FA for 16 h. (<b>E</b>) The fluorescence intensities of both nanoparticles in blood over time. The error bars indicate the mean ± SD (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span>&lt; 0.01. Reproduced with permission from [<a href="#B39-nanomaterials-06-00081" class="html-bibr">39</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 5
<p>Experimental design for the large stokes shifting NIR fluorescent SiNPs (LSS-NFSiNPs) and real-time abdomen fluorescence resonance energy transfer (FRET) imaging of mice intravenously injected with the LSS-NFSiNPs. Reproduced with permission from [<a href="#B30-nanomaterials-06-00081" class="html-bibr">30</a>]. Copyright 2012 American Chemical Society.</p>
Full article ">Figure 6
<p>Schematic illustration of dye-loaded mesoporous SiNPs for both NIR fluorescent and photoacoustic (PA) imaging. Reproduced with permission from [<a href="#B72-nanomaterials-06-00081" class="html-bibr">72</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 7
<p>Schematic illustration of organically-modified two-photon photodynamic therapy SiNPs. Reproduced with permission from [<a href="#B74-nanomaterials-06-00081" class="html-bibr">74</a>]. Copyright 2007 American Chemical Society.</p>
Full article ">Figure 8
<p>Schematic illustration of two-photon (TP)-MSNs@MnO<sub>2</sub>for glutathione detection. Reproduced with permission from [<a href="#B73-nanomaterials-06-00081" class="html-bibr">73</a>]. Copyright 2014 American Chemical Society.</p>
Full article ">Figure 9
<p>Two-photon confocal microscopy (CM) images of glutathione detection in living CEM cells. (<b>a</b>) CEM cells incubated with the TP-MSN@MnO<sub>2</sub> nanoparticle; (<b>b</b>) CEM cells pretreated with glutathione scavenger and then incubated with the TP-MSN@MnO<sub>2</sub> nanoparticle; (<b>c</b>) CEM cells pretreated with glutathione synthesis enhancer and incubated with the TP-MSN@MnO<sub>2</sub> nanoparticle. Reproduced with permission from [<a href="#B73-nanomaterials-06-00081" class="html-bibr">73</a>]. Copyright 2014 American Chemical Society.</p>
Full article ">Figure 10
<p>Schematic illustration of MSN probe-based intracellular detection of telomerase. Reproduced with permission from [<a href="#B76-nanomaterials-06-00081" class="html-bibr">76</a>]. Copyright 2013 American Chemical Society. BHQ, black hole fluorescence quencher.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop