[go: up one dir, main page]

Next Issue
Volume 6, February
Previous Issue
Volume 5, December
 
 

Nanomaterials, Volume 6, Issue 1 (January 2016) – 23 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:
171 KiB  
Editorial
Acknowledgement to Reviewers of Nanomaterials in 2015
by Nanomaterials Editorial Office
Nanomaterials 2016, 6(1), 23; https://doi.org/10.3390/nano6010023 - 21 Jan 2016
Viewed by 3696
Abstract
The editors of Nanomaterials would like to express their sincere gratitude to the following reviewers for assessing manuscripts in 2015. [...] Full article
5191 KiB  
Article
Coupling of Nanocrystalline Anatase TiO2 to Porous Nanosized LaFeO3 for Efficient Visible-Light Photocatalytic Degradation of Pollutants
by Muhammad Humayun, Zhijun Li, Liqun Sun, Xuliang Zhang, Fazal Raziq, Amir Zada, Yang Qu and Liqiang Jing
Nanomaterials 2016, 6(1), 22; https://doi.org/10.3390/nano6010022 - 20 Jan 2016
Cited by 37 | Viewed by 7293
Abstract
In this work we have successfully fabricated nanocrystalline anatase TiO2/perovskite-type porous nanosized LaFeO3 (T/P-LFO) nanocomposites using a simple wet chemical method. It is clearly demonstrated by means of atmosphere-controlled steady-state surface photovoltage spectroscopy (SPS) responses, photoluminescence spectra, and fluorescence spectra [...] Read more.
In this work we have successfully fabricated nanocrystalline anatase TiO2/perovskite-type porous nanosized LaFeO3 (T/P-LFO) nanocomposites using a simple wet chemical method. It is clearly demonstrated by means of atmosphere-controlled steady-state surface photovoltage spectroscopy (SPS) responses, photoluminescence spectra, and fluorescence spectra related to the formed OH radical amount that the photogenerated charge carriers in the resultant T/P-LFO nanocomposites with a proper mole ratio percentage of TiO2 display much higher separation in comparison to the P-LFO alone. This is highly responsible for the improved visible-light activities of T/P-LFO nanocomposites for photocatalytic degradation of gas-phase acetaldehyde and liquid-phase phenol. This work will provide a feasible route to synthesize visible-light responsive nano-photocatalysts for efficient solar energy utilization. Full article
(This article belongs to the Special Issue Nanoparticles for Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffraction (XRD) patterns (<b>A</b>) and UV-vis diffuse reflectance (UV-vis DRS) spectra (<b>B</b>) of Porous-LaFeO<sub>3</sub> (P-LFO) TiO<sub>2</sub> (T) and TiO<sub>2</sub>/P-LFO (T/P-LFO) nanocomposites; Transmission electron microscopy (TEM) image of P-LFO with inset selected area electron diffraction (SAED) pattern (<b>C</b>); TEM image of 9T/P-LFO with inset high resolution transmission electron microscopy (HRTEM) image (<b>D</b>).</p>
Full article ">Figure 2
<p>Scanning electron microscopy (SEM) micrograph of P-LFO (<b>A</b>) and 9T/P-LFO nanocomposite (<b>B</b>).</p>
Full article ">Figure 3
<p>X-ray photoelectron spectroscopy (XPS) survey spectra of P-LFO and 9T/P-LFO nanocomposite (<b>A</b>); with high resolution images La3d (<b>B</b>); Fe2p (<b>C</b>); O1s (<b>D</b>); Ti2p (<b>E</b>).</p>
Full article ">Figure 4
<p>N<sub>2</sub> adsorption/desorption isotherms (<b>A</b>) and pore diameter (<b>B</b>) of P-LFO and 9T/P-LFO nanocomposite.</p>
Full article ">Figure 5
<p>Surface photovoltage spectroscopy (SPS) responses of P-LFO (<b>A</b>) and 9T/P-LFO (<b>B</b>) in different atmospheres; SPS responses of P-LFO and T/P-LFO nanocomposites in air (<b>C</b>); Photoluminescence (PL) responses of P-LFO and T/P-LFO nanocomposites (<b>D</b>).</p>
Full article ">Figure 6
<p>Visible-light photocatalytic activity for acetaldehyde and phenol degradation (<b>A</b>) and OH<sup>−</sup> radical amount related Fluorescence spectra (<b>B</b>) of P-LFO and T/P-LFO nanocomposites.</p>
Full article ">Figure 7
<p>Scheme for energy band gaps and the mechanism for photogenerated charge separation and transfer in the fabricated T/P-LFO nanocomposite.</p>
Full article ">
2253 KiB  
Article
Investigation of MnO2 and Ordered Mesoporous Carbon Composites as Electrocatalysts for Li-O2 Battery Applications
by Chih-Chun Chin, Hong-Kai Yang and Jenn-Shing Chen
Nanomaterials 2016, 6(1), 21; https://doi.org/10.3390/nano6010021 - 18 Jan 2016
Cited by 16 | Viewed by 8485
Abstract
The electrocatalytic activities of the MnO2/C composites are examined in Li-O2 cells as the cathode catalysts. Hierarchically mesoporous carbon-supported manganese oxide (MnO2/C) composites are prepared using a combination of soft template and hydrothermal methods. The composites are characterized [...] Read more.
The electrocatalytic activities of the MnO2/C composites are examined in Li-O2 cells as the cathode catalysts. Hierarchically mesoporous carbon-supported manganese oxide (MnO2/C) composites are prepared using a combination of soft template and hydrothermal methods. The composites are characterized by X-ray powder diffraction, scanning electron microscopy, transmission electron microscopy, small angle X-ray scattering, The Brunauer–Emmett–Teller (BET) measurements, galvanostatic charge-discharge methods, and rotating ring-disk electrode (RRDE) measurements. The electrochemical tests indicate that the MnO2/C composites have excellent catalytic activity towards oxygen reduction reactions (ORRs) due to the larger surface area of ordered mesoporous carbon and higher catalytic activity of MnO2. The O2 solubility, diffusion rates of O2 and O2•− coefficients (DO2 and DO2), the rate constant (kf) for producing O2•−, and the propylene carbonate (PC)-electrolyte
decomposition rate constant (k) of the MnO2/C material were measured by RRDE experiments in the 0.1 M TBAPF6/PC electrolyte. The values of kf and k for MnO2/C are 4.29 × 10−2 cm·s−1 and 2.6 s−1, respectively. The results indicate that the MnO2/C cathode catalyst has higher electrocatalytic activity for the first step of ORR to produce O2•− and achieves a faster PC-electrolyte decomposition rate. Full article
(This article belongs to the Special Issue Nanoparticles for Catalysis)
Show Figures

Figure 1

Figure 1
<p>Schematic of a four-neck, jacketed glass cell with a rotating ring-disk electrode (RRDE) system.</p>
Full article ">Figure 2
<p>(<b>a</b>) Wide-angle X-ray diffraction (XRD) patterns; (<b>b</b>) small angle X-ray scattering (SAXS) patterns of MnO<sub>2</sub>/C composites.</p>
Full article ">Figure 3
<p>Scanning electron microscope (SEM) images (<b>a</b>) MnO<sub>2</sub>/C composites; (<b>b</b>) high magnification of the region marked with a square in (<b>a</b>); and transmission electron microscope (TEM) images (<b>c</b>,<b>e</b>) MnO<sub>2</sub>/C composites; (<b>d</b>) high magnification of the region marked with a square in (<b>c</b>); and (<b>f</b>) selected area electron diffraction (SAED) pattern of the region marked with a square in (<b>e</b>).</p>
Full article ">Figure 4
<p>Nitrogen sorption isotherms of MnO<sub>2</sub>/C composites. The insert is the Barrett–Joyner–Halenda (BJH) desorption pore size distribution.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curves were recorded at a scanning rate of 2 mV·s<sup>−1</sup> for MnO<sub>2</sub>/C and Super-P carbon samples; (<b>b</b>) initial charge–discharge profiles for MnO<sub>2</sub>/C and Super P samples at a current density of 0.2 mA·cm<sup>−2</sup>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Example of determination of the superoxide radical (O<sub>2</sub><sup>•−</sup>) transit-time (<span class="html-italic">T<sub>s</sub></span>) in O<sub>2</sub>-saturated solutions of 0.1 M TBAPF<sub>6</sub> in propylene carbonate (PC) at ω = 100 rpm, E<sub>disk</sub> = 1.85 V and E<sub>ring</sub> = 2.6 V. Transit time (<span class="html-italic">Ts</span>) values at different rotation rates for the diffusion of (<b>b</b>) O<sub>2</sub> and (<b>c</b>) O<sub>2</sub><sup>•−</sup>; (<b>d</b>) relation between the inverse of the rotation speed and the transient time for O<sub>2</sub> and O<sub>2</sub><sup>•−</sup>.</p>
Full article ">Figure 7
<p>(<b>a</b>) Steady-state CV curves of a glassy carbon rotating disk electrode (RDE) in an O<sub>2</sub>-saturated 0.1 M TBAPF<sub>6</sub>/PC solution at a scan rate of 50 mV/s between 1.5 and 2.8 V<sub>Li</sub> with different rotation rates. The insert is the Koutecky–Levich plot derived from the disc current values at 1.50 V<sub>Li</sub>; (<b>b</b>) steady-state CV curves of a MnO<sub>2</sub>/C RDE in an O<sub>2</sub>-saturated 0.1 M TBAPF<sub>6</sub>/PC solution at a scan rate of 50 mV/s between 1.2 and 2.8 V<sub>Li</sub> with different rotation rates.</p>
Full article ">Figure 8
<p>(<b>a</b>) RRDE profiles of MnO<sub>2</sub>/C recorded at 50 mV·s<sup>−1</sup> in an O<sub>2</sub>-saturated 0.1 M TBAPF<sub>6</sub>/PC solution, at rotation rates between 300 and 2100 rpm with continuous holding of the Pt ring at 2.85 V<sub>Li</sub>; (<b>b</b>) evolution of the absolute ratio between the ring and disk current (<span class="html-italic">N<sub>k</sub></span>) and the electrode rotation speed (ω).</p>
Full article ">
1396 KiB  
Communication
A Graphene Oxide-Based Fluorescent Platform for Probing of Phosphatase Activity
by Ting Sun, Ning Xia and Lin Liu
Nanomaterials 2016, 6(1), 20; https://doi.org/10.3390/nano6010020 - 18 Jan 2016
Cited by 13 | Viewed by 5872
Abstract
We presented a strategy for fabricating graphene oxide (GO)-based fluorescent biosensors to monitor the change of phosphorylation state and detect phosphatase activity. By regulating the interaction between the negatively charged phosphate group and the positively charged amino residue, we found that GO showed [...] Read more.
We presented a strategy for fabricating graphene oxide (GO)-based fluorescent biosensors to monitor the change of phosphorylation state and detect phosphatase activity. By regulating the interaction between the negatively charged phosphate group and the positively charged amino residue, we found that GO showed different quenching efficiency toward the phosphorylated and dephosphorylated dye-labeled peptides. To demonstrate the application of our method, alkaline phosphatase (ALP) was tested as a model enzyme with phosphorylated fluorescein isothiocyanate (FITC)-labeled short peptide FITC–Gly–Gly–Gly–Tyr(PO32−)–Arg as the probe. When the negatively charged phosphate group in the Tyr residue was removed from the peptide substrate by enzymatic hydrolysis, the resulting FITC–Gly–Gly–Gly–Tyr–Arg was readily adsorbed onto the GO surface through electrostatic interaction. As a result, fluorescence quenching was observed. Furthermore, the method was applied for the screening of phosphatase inhibitors. Full article
(This article belongs to the Special Issue Nanomaterials for Biosensing Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of dephosphorylation process (<b>A</b>) and alkaline phosphatase (ALP) activity detection (<b>B</b>) using fluorescein isothiocyanate (FITC)–Gly–Gly–Gly–Tyr(PO<sub>3</sub><sup>2−</sup>)–Arg (GGGYpR) and graphene oxide (GO) as the probe and the quencher, respectively.</p>
Full article ">Figure 2
<p>Quenching efficiency of different concentrations of GO toward phosphorylated peptide FITC–GGGYpR and dephosphorylated peptide FITC–Gly–Gly–Gly–Tyr–Arg (GGGYR).</p>
Full article ">Figure 3
<p>(<b>A</b>) Fluorescence spectra of FITC–GGGYpR in the presence of GO and ALP; (<b>B</b>) Fluorescence intensity of FITC–GGGYpR/GO in the absence (bar 1) and presence of ALP (bar 2), bovine serum albumin (BSA) (bar 3), lysozyme (bar 4), myoglobin (bar 5), and ALP/BSA/lysozyme/myoglobin (bar 6). The final concentrations of FITC–GGGYpR, GO, ALP, BSA, lysozyme and myoglobin are 100 nM, 50 μg·mL<sup>−1</sup>, 10 nM, 0.5 μg·mL<sup>−1</sup>, 0.5 μg·mL<sup>−1</sup> and 0.5 μg·mL<sup>−1</sup>, respectively.</p>
Full article ">Figure 4
<p>(<b>A</b>) Fluorescence intensity of FITC–GGGYpR/GO after addition of different concentrations of ALP (from a to j: 0, 0.2, 0.5, 1, 2, 3, 4, 5, 7.5 and 10 nM). The final concentrations of GO and FITC–GGGYpR are 50 μg∙mL<sup>−1</sup> and 100 nM, respectively; (<b>B</b>) Dependence of fluorescence intensity on the ALP concentration.</p>
Full article ">Figure 5
<p>(<b>A</b>) Fluorescence spectra of 100 nM FITC–GGGYpR/GO in the presence of 10 nM ALP and different concentrations of levamisole (0, 5, 10, 20, 40, 60, 80, 100 and 120 nM); (<b>B</b>) Dependence of the fluorescence intensity on the concentration of levamisole.</p>
Full article ">
3324 KiB  
Article
Synthesis of Nickel Nanowires with Tunable Characteristics
by Zengzilu Xia and Weijia Wen
Nanomaterials 2016, 6(1), 19; https://doi.org/10.3390/nano6010019 - 15 Jan 2016
Cited by 27 | Viewed by 8336
Abstract
A one-step synthesis of magnetic nickel nanowires (NiNWs) with tunable characteristics is reported. The method is simple and easy to be conducted, leading to high compatibility with scaling-up. It is discovered that the size and morphology of NiNWs can be adjusted by tuning [...] Read more.
A one-step synthesis of magnetic nickel nanowires (NiNWs) with tunable characteristics is reported. The method is simple and easy to be conducted, leading to high compatibility with scaling-up. It is discovered that the size and morphology of NiNWs can be adjusted by tuning the reaction temperature, time length, as well as surfactant concentration. It is found that the products have shown high purity which remained after being stored for several months. A remarkable enhanced saturation magnetization of the product was also observed, compared to that of bulk nickel. By providing both practical experimental details and in-depth mechanism, the work introduced in this paper may advance the mass production and further applications of NiNWs. Full article
(This article belongs to the Special Issue Nanoparticles in Bioimaging)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The photographs of nickel nanowires (NiNWs) (<b>a</b>) dispersed in ethanol (<b>b</b>) attracted by an external magnet, (<b>c</b>) X-ray diffraction (XRD) patterns, (<b>d</b>) low and (<b>e</b>) high magnification scanning electron microscopy (SEM) images of NiNWs; and the transmission electron microscopy (TEM) images of (<b>f</b>) the body and (<b>g</b>) the end of a NiNW.</p>
Full article ">Figure 1 Cont.
<p>The photographs of nickel nanowires (NiNWs) (<b>a</b>) dispersed in ethanol (<b>b</b>) attracted by an external magnet, (<b>c</b>) X-ray diffraction (XRD) patterns, (<b>d</b>) low and (<b>e</b>) high magnification scanning electron microscopy (SEM) images of NiNWs; and the transmission electron microscopy (TEM) images of (<b>f</b>) the body and (<b>g</b>) the end of a NiNW.</p>
Full article ">Figure 2
<p>SEM images with same magnification of NiNWs synthesized at (<b>a</b>) 70 °C, (<b>b</b>) 110 °C, and (<b>c</b>) 150 °C. The change of (<b>d</b>) length, (<b>e</b>) width, and (<b>f</b>) length-to-width ratio (LWR) of NiNWs with temperature.</p>
Full article ">Figure 3
<p>Typical SEM images with same magnification of NiNWs synthesized with (<b>a</b>) 1 min, (<b>b</b>) 10 min, and (<b>c</b>) 90 min. The change of (<b>d</b>) length, (<b>e</b>) width, and (<b>f</b>) LWR of NiNWs with reaction time.</p>
Full article ">Figure 4
<p>SEM images with same magnification of NiNWs synthesized with poly(vinylpyrrolidone) (PVP) concentration of (<b>a</b>) 0.03 <span class="html-italic">w</span>/<span class="html-italic">v</span> %, (<b>b</b>) 0.5 <span class="html-italic">w</span>/<span class="html-italic">v</span> %, and (<b>c</b>) 2 <span class="html-italic">w</span>/<span class="html-italic">v</span> %. The change of (<b>d</b>) length, (<b>e</b>) width, and (<b>f</b>) LWR of NiNWs with PVP concentration.</p>
Full article ">Figure 5
<p>The typical energy-dispersive X-ray spectroscopy (EDS) spectrum. The inset is hysteretic loop at room temperature of typical NiNWs. These characterizations were conducted on the product synthesized with typical process.</p>
Full article ">Figure 6
<p>The illustration of NiNW growth mechanism. The inset is the detailed illustration of Step 2 to 4.</p>
Full article ">
3758 KiB  
Communication
Nitrogen-Doped Banana Peel–Derived Porous Carbon Foam as Binder-Free Electrode for Supercapacitors
by Bingzhi Liu, Lili Zhang, Peirong Qi, Mingyuan Zhu, Gang Wang, Yanqing Ma, Xuhong Guo, Hui Chen, Boya Zhang, Zhuangzhi Zhao, Bin Dai and Feng Yu
Nanomaterials 2016, 6(1), 18; https://doi.org/10.3390/nano6010018 - 15 Jan 2016
Cited by 63 | Viewed by 10746
Abstract
Nitrogen-doped banana peel–derived porous carbon foam (N-BPPCF) successfully prepared from banana peels is used as a binder-free electrode for supercapacitors. The N-BPPCF exhibits superior performance including high specific surface areas of 1357.6 m2/g, large pore volume of 0.77 cm3/g, [...] Read more.
Nitrogen-doped banana peel–derived porous carbon foam (N-BPPCF) successfully prepared from banana peels is used as a binder-free electrode for supercapacitors. The N-BPPCF exhibits superior performance including high specific surface areas of 1357.6 m2/g, large pore volume of 0.77 cm3/g, suitable mesopore size distributions around 3.9 nm, and super hydrophilicity with nitrogen-containing functional groups. It can easily be brought into contact with an electrolyte to facilitate electron and ion diffusion. A comparative analysis on the electrochemical properties of BPPCF electrodes is also conducted under similar conditions. The N-BPPCF electrode offers high specific capacitance of 185.8 F/g at 5 mV/s and 210.6 F/g at 0.5 A/g in 6 M KOH aqueous electrolyte versus 125.5 F/g at 5 mV/s and 173.1 F/g at 0.5 A/g for the BPPCF electrode. The results indicate that the N-BPPCF is a binder-free electrode that can be used for high performance supercapacitors. Full article
(This article belongs to the Special Issue Nanostructured Materials for Li-Ion Batteries and Beyond)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Synthesis of porous carbon foam from banana peels. Scanning electron microscopy (SEM) images of (<b>b</b>) banana peel-derived porous carbon foam (BPPCF) and (<b>c</b>) nitrogen-doped BPPCF (N-BPPCF).</p>
Full article ">Figure 2
<p>Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images at different magnifications of (<b>a</b>, <b>c</b> and <b>e</b>) BPPCF and (<b>b</b>, <b>d</b> and <b>f</b>) N-BPPCF.</p>
Full article ">Figure 3
<p>Nitrogen sorption isotherms of (<b>a</b>) BPPCF and (<b>b</b>) N-BPPCF, and the corresponding pore size distribution (insets).</p>
Full article ">Figure 4
<p>(<b>a</b>) X-ray diffraction (XRD) patterns for BPPCF and N-BPPCF samples. X-ray photoelectron spectroscopy (XPS) spectra of (<b>b</b>) C 1s, (<b>c</b>) O 1s and (<b>d</b>) N 1s for as-obtained BPPCF and N-BPPCF.</p>
Full article ">Figure 5
<p>Cyclic voltammogram (CV) curves for (<b>a</b>) BPPCF and (<b>b</b>) N-BPPCF in 6 M KOH at different scan rates. (<b>c</b>) CV curves of the samples at a scan rate of 5 mV/s. (<b>d</b>) Specific capacitance calculated according to CVs.</p>
Full article ">Figure 6
<p>Galvanostatic charge/discharge (GCD) curves for (<b>a</b>) BPPCF and (<b>b</b>) N-BPPCF in 6 M KOH at various current densities. (<b>c</b>) Specific capacitance and (<b>d</b>) the corresponding rate capacitance retention of BPPCF and N-BPPCF.</p>
Full article ">Figure 7
<p>Cycle life of BPPCF and N-BPPCF at various currents: (<b>a</b>) 0.5 A/g for 500 cycles and (<b>b</b>) 2.5 A/g for 5000 cycles.</p>
Full article ">
2269 KiB  
Review
Excipient Nanoemulsions for Improving Oral Bioavailability of Bioactives
by Laura Salvia-Trujillo, Olga Martín-Belloso and David Julian McClements
Nanomaterials 2016, 6(1), 17; https://doi.org/10.3390/nano6010017 - 14 Jan 2016
Cited by 93 | Viewed by 12718
Abstract
The oral bioavailability of many hydrophobic bioactive compounds found in natural food products (such as vitamins and nutraceuticals in fruits and vegetables) is relatively low due to their low bioaccessibility, chemical instability, or poor absorption. Most previous research has therefore focused on the [...] Read more.
The oral bioavailability of many hydrophobic bioactive compounds found in natural food products (such as vitamins and nutraceuticals in fruits and vegetables) is relatively low due to their low bioaccessibility, chemical instability, or poor absorption. Most previous research has therefore focused on the design of delivery systems to incorporate isolated bioactive compounds into food products. However, a more sustainable and cost-effect approach to enhancing the functionality of bioactive compounds is to leave them within their natural environment, but specifically design excipient foods that enhance their bioavailability. Excipient foods typically do not have functionality themselves but they have the capacity to enhance the functionality of nutrients present in natural foods by altering their bioaccessibility, absorption, and/or chemical transformation. In this review article we present the use of excipient nanoemulsions for increasing the bioavailability of bioactive components from fruits and vegetables. Nanoemulsions present several advantages over other food systems for this application, such as the ability to incorporate hydrophilic, amphiphilic, and lipophilic excipient ingredients, high physical stability, and rapid gastrointestinal digestibility. The design, fabrication, and application of nanoemulsions as excipient foods will therefore be described in this article. Full article
(This article belongs to the Special Issue Nanoparticles Assisted Drug Delivery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the difference between integrated and non-integrated excipient foods. For integrated excipient foods the bioactive component (pharmaceutical or nutraceutical) is dispersed within the excipient food matrix, but for non-integrated excipient foods the bioactive component is in another product that is co-ingested with the excipient food.</p>
Full article ">Figure 2
<p>The overall oral bioavailability of bioactives is governed by three main factors: bioaccessibility; absorption; and transformation.</p>
Full article ">Figure 3
<p>Schematic representation of the variables that can be modulated in order to formulate excipient nanoemulsions to enhance the oral bioavailability of naturally-occurring bioactive compounds.</p>
Full article ">Figure 4
<p>Schematic representation of the mechanisms involved in the absorption of bioactive compounds in the gastrointestinal tract. The absorption of bioactive agents may be limited due to their transport across the epithelium cell through passive, active or efflux mechanisms. Most bioactive compounds are usually absorbed in the upper gastrointestinal tract (GIT) and therefore do not reach the M-cells, but bioactives trapped within indigestible particles or matrices may move further down the GIT and then encounter the M-cells.</p>
Full article ">
3001 KiB  
Article
Resistive Switching of Plasma–Treated Zinc Oxide Nanowires for Resistive Random Access Memory
by Yunfeng Lai, Wenbiao Qiu, Zecun Zeng, Shuying Cheng, Jinling Yu and Qiao Zheng
Nanomaterials 2016, 6(1), 16; https://doi.org/10.3390/nano6010016 - 13 Jan 2016
Cited by 23 | Viewed by 6850
Abstract
ZnO nanowires (NWs) were grown on Si(100) substrates at 975 °C by a vapor-liquid-solid method with ~2 nm and ~4 nm gold thin films as catalysts, followed by an argon plasma treatment for the as-grown ZnO NWs. A single ZnO NW–based memory cell [...] Read more.
ZnO nanowires (NWs) were grown on Si(100) substrates at 975 °C by a vapor-liquid-solid method with ~2 nm and ~4 nm gold thin films as catalysts, followed by an argon plasma treatment for the as-grown ZnO NWs. A single ZnO NW–based memory cell with a Ti/ZnO/Ti structure was then fabricated to investigate the effects of plasma treatment on the resistive switching. The plasma treatment improves the homogeneity and reproducibility of the resistive switching of the ZnO NWs, and it also reduces the switching (set and reset) voltages with less fluctuations, which would be associated with the increased density of oxygen vacancies to facilitate the resistive switching as well as to average out the stochastic movement of individual oxygen vacancies. Additionally, a single ZnO NW–based memory cell with self-rectification could also be obtained, if the inhomogeneous plasma treatment is applied to the two Ti/ZnO contacts. The plasma-induced oxygen vacancy disabling the rectification capability at one of the Ti/ZnO contacts is believed to be responsible for the self-rectification in the memory cell. Full article
(This article belongs to the Special Issue Plasma Nanoengineering and Nanofabrication)
Show Figures

Figure 1

Figure 1
<p>Reproducible voltage-biased current-voltage (<span class="html-italic">I</span>–<span class="html-italic">V</span>) curves of the ZnO nanowires (ZnO NW)–based memories (<b>a</b>,<b>c</b>) without and (<b>b</b>,<b>d</b>) with argon plasma treatment. Blue (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="blue"> <mo>☆</mo> </mstyle> </mrow> </semantics> </math>), black (<math display="inline"> <semantics> <mrow> <mstyle mathvariant="bold" mathsize="normal"> <mo>○</mo> </mstyle> </mrow> </semantics> </math>) and red (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>△</mo> </mstyle> </mrow> </semantics> </math>) symbols respectively represent the 8th, 18th and 28th switching cycles.</p>
Full article ">Figure 2
<p>Temperature-dependent resistance of the low resistance state (LRS) memories with and without plasma treatment.</p>
Full article ">Figure 3
<p>Endurance tests of the ZnO NW–based memories (<b>a</b>) without and (<b>b</b>) with argon plasma treatment.</p>
Full article ">Figure 4
<p>Distributions of switching voltages of the ZnO NW–based memories (<b>a</b>) without and (<b>b</b>) with argon plasma treatment.</p>
Full article ">Figure 5
<p>Data retention of the ZnO NW–based memories (<b>a</b>) without and (<b>b</b>) with argon plasma treatment.</p>
Full article ">Figure 6
<p>(<b>a</b>) Reproducible asymmetric <span class="html-italic">I</span>–<span class="html-italic">V</span> curves of single ZnO NW and (<b>b</b>) the <span class="html-italic">I</span>–<span class="html-italic">V</span> curves at positive bias on log-log scale with the inset fitting of <math display="inline"> <semantics> <mrow> <msqrt> <mi>V</mi> </msqrt> <mo>~</mo> <mi>ln</mi> <mo stretchy="false">(</mo> <mi>I</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> for the high resistance state (HRS).</p>
Full article ">Figure 7
<p>Cross-sectional scanning electron microscope (SEM) images of the ZnO NWs on the silicon substrates with gold thicknesses of (<b>a</b>) ~2 nm and (<b>b</b>) ~4 nm.</p>
Full article ">Figure 8
<p>Room-temperature photoluminescence (PL) spectra of the ZnO NWs and their Gaussian components.</p>
Full article ">Figure 9
<p>Top-view SEM images of the single ZnO NW–based resistive random access memory (RRAM) cell.</p>
Full article ">
157 KiB  
Editorial
Frontiers in Mesoporous Nanomaterials
by Eva Pellicer and Jordi Sort
Nanomaterials 2016, 6(1), 15; https://doi.org/10.3390/nano6010015 - 11 Jan 2016
Cited by 3 | Viewed by 4254
Abstract
The Special Issue of Nanomaterials “Frontiers in Mesoporous Nanomaterials” gathers four reviews, one communication and eight regular papers. Full article
(This article belongs to the Special Issue Frontiers in Mesoporous Nanomaterials)
979 KiB  
Article
X-ray Absorption Spectroscopy Characterization of a Li/S Cell
by Yifan Ye, Ayako Kawase, Min-Kyu Song, Bingmei Feng, Yi-Sheng Liu, Matthew A. Marcus, Jun Feng, Elton J. Cairns, Jinghua Guo and Junfa Zhu
Nanomaterials 2016, 6(1), 14; https://doi.org/10.3390/nano6010014 - 11 Jan 2016
Cited by 30 | Viewed by 9868
Abstract
The X-ray absorption spectroscopy technique has been applied to study different stages of the lithium/sulfur (Li/S) cell life cycle. We have investigated how speciation of S in Li/S cathodes changes upon the introduction of CTAB (cetyltrimethylammonium bromide, CH3(CH2)15 [...] Read more.
The X-ray absorption spectroscopy technique has been applied to study different stages of the lithium/sulfur (Li/S) cell life cycle. We have investigated how speciation of S in Li/S cathodes changes upon the introduction of CTAB (cetyltrimethylammonium bromide, CH3(CH2)15N+(CH3)3Br) and with charge/discharge cycling. The introduction of CTAB changes the synthesis reaction pathway dramatically due to the interaction of CTAB with the terminal S atoms of the polysulfide ions in the Na2Sx solution. For the cycled Li/S cell, the loss of electrochemically active sulfur and the accumulation of a compact blocking insulating layer of unexpected sulfur reaction products on the cathode surface during the charge/discharge processes make the capacity decay. A modified coin cell and a vacuum-compatible three-electrode electro-chemical cell have been introduced for further in-situ/in-operando studies. Full article
(This article belongs to the Special Issue Nanostructured Materials for Li-Ion Batteries and Beyond)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>S K-edge X-ray absorption spectroscopy (XAS) data of (<b>a</b>) Na<sub>2</sub>S<span class="html-italic"><sub>X</sub></span> and Na<sub>2</sub>S<span class="html-italic"><sub>X</sub></span> + cetyltrimethylammonium bromide, CH<sub>3</sub>(CH<sub>2</sub>)<sub>15</sub>N<sup>+</sup>(CH<sub>3</sub>)<sub>3</sub>Br<sup>−</sup> (CTAB) solutions, (<b>b</b>) the precipitates collected from the acidified solutions. The fitting of the lower energy region of the spectra of (<b>a</b>) is shown in (<b>c</b>), while the intensities of peaks A and B and the ratio of Peak B/Peak A are shown in (<b>d</b>).</p>
Full article ">Figure 2
<p>The (<b>a</b>) S K-edge XAS and (<b>b</b>) C K-edge XAS data of cathode materials cycled for 0, 500 and 1500 cycles.</p>
Full article ">Figure 3
<p>The scheme of (<b>a</b>) a modified coin cell and (<b>b</b>) a three-electrode electro-chemical cell (figure adapted from Reference [<a href="#B14-nanomaterials-06-00014" class="html-bibr">14</a>]); and X-ray transmission at (<b>c</b>) soft X-ray energy region and (<b>d</b>) tender X-ray energy region through different materials.</p>
Full article ">
1724 KiB  
Article
White Light-Emitting Diodes Based on AgInS2/ZnS Quantum Dots with Improved Bandwidth in Visible Light Communication
by Cheng Ruan, Yu Zhang, Min Lu, Changyin Ji, Chun Sun, Xiongbin Chen, Hongda Chen, Vicki L. Colvin and William W. Yu
Nanomaterials 2016, 6(1), 13; https://doi.org/10.3390/nano6010013 - 8 Jan 2016
Cited by 55 | Viewed by 10692
Abstract
Quantum dot white light-emitting diodes (QD-WLEDs) were fabricated from green- and red-emitting AgInS2/ZnS core/shell QDs coated on GaN LEDs. Their electroluminescence (EL) spectra were measured at different currents, ranging from 50 mA to 400 mA, and showed good color stability. The [...] Read more.
Quantum dot white light-emitting diodes (QD-WLEDs) were fabricated from green- and red-emitting AgInS2/ZnS core/shell QDs coated on GaN LEDs. Their electroluminescence (EL) spectra were measured at different currents, ranging from 50 mA to 400 mA, and showed good color stability. The modulation bandwidth of previously prepared QD-WLEDs was confirmed to be much wider than that of YAG:Ce phosphor-based WLEDs. These results indicate that the AgInS2/ZnS core/shell QDs are good color-converting materials for WLEDs and they are capable in visible light communication (VLC). Full article
(This article belongs to the Special Issue Current Trends in Colloidal Nanocrystals)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Absorption spectra and (<b>b</b>) photoluminescence (PL) spectra of green- and red-emitting AgInS<sub>2</sub>/ZnS Quantum dots (QDs) in hexane. The inset shows the corresponding real color under excitation of 365 nm UV light.</p>
Full article ">Figure 2
<p>(<b>a</b>) The device structure for generating white light from green- and red-emitting AgInS<sub>2</sub>/ZnS QDs; (<b>b</b>) the real emitting color picture operated at 350 mA; (<b>c</b>) electroluminescence (EL spectra) of AgInS<sub>2</sub> Quantum dot white light-emitting diode (QD-WLED) at different working currents from 50 mA to 400 mA; (<b>d</b>) the corresponding CIE coordinates of the QD-WLED.</p>
Full article ">Figure 3
<p>The peak positions and the full width at half-maximum (FWHM) of QDs in the white light emitting diode (WLED) under different currents from 50 mA to 400 mA for (<b>a</b>) green- and (<b>b</b>) red-emitting AgInS<sub>2</sub>/ZnS QDs; (<b>c</b>) PL spectra of the WLED at different working time when the working current was 350 mA.</p>
Full article ">Figure 4
<p>(<b>a</b>) The frequency responses of blue GaN LED (black line), AgInS<sub>2</sub>/ZnS QD-WLED (red line), and YAG:Ce WLED (blue line) at 350 mA, respectively; (<b>b</b>) PL decay curves of green- and red-emitting AgInS<sub>2</sub>/ZnS QDs.</p>
Full article ">
947 KiB  
Article
The Coupled Photothermal Reaction and Transport in a Laser Additive Metal Nanolayer Simultaneous Synthesis and Pattering for Flexible Electronics
by Song-Ling Tsai, Yi-Kai Liu, Heng Pan, Chien-Hung Liu and Ming-Tsang Lee
Nanomaterials 2016, 6(1), 12; https://doi.org/10.3390/nano6010012 - 8 Jan 2016
Cited by 8 | Viewed by 6344
Abstract
The Laser Direct Synthesis and Patterning (LDSP) technology has advantages in terms of processing time and cost compared to nanomaterials-based laser additive microfabrication processes. In LDSP, a scanning laser on the substrate surface induces chemical reactions in the reactive liquid solution and selectively [...] Read more.
The Laser Direct Synthesis and Patterning (LDSP) technology has advantages in terms of processing time and cost compared to nanomaterials-based laser additive microfabrication processes. In LDSP, a scanning laser on the substrate surface induces chemical reactions in the reactive liquid solution and selectively deposits target material in a preselected pattern on the substrate. In this study, we experimentally investigated the effect of the processing parameters and type and concentration of the additive solvent on the properties and growth rate of the resulting metal film fabricated by this LDSP technology. It was shown that reactive metal ion solutions with substantial viscosity yield metal films with superior physical properties. A numerical analysis was also carried out the first time to investigate the coupled opto-thermo-fluidic transport phenomena and the effects on the metal film growth rate. To complete the simulation, the optical properties of the LDSP deposited metal film with a variety of thicknesses were measured. The characteristics of the temperature field and the thermally induced flow associated with the moving heat source are discussed. It was shown that the processing temperature range of the LDSP is from 330 to 390 K. A semi-empirical model for estimating the metal film growth rate using this process was developed based on these results. From the experimental and numerical results, it is seen that, owing to the increased reflectivity of the silver film as its thickness increases, the growth rate decreases gradually from about 40 nm at initial to 10 nm per laser scan after ten scans. This self-controlling effect of LDSP process controls the thickness and improves the uniformity of the fabricated metal film. The growth rate and resulting thickness of the metal film can also be regulated by adjustment of the processing parameters, and thus can be utilized for controllable additive nano/microfabrication. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The silver ion reaction solution used in this study; and (<b>b</b>) The absorbance spectrum of the silver ion solution and its mixture (1:1) with ethylene glycol or propylene glycol.</p>
Full article ">Figure 2
<p>Experimental apparatus of laser direct synthesis and patterning (LDSP).</p>
Full article ">Figure 3
<p>A schematic of the simulation domain used in the current study.</p>
Full article ">Figure 4
<p>SEM images of the silver film fabricated from the reaction solution with (<b>a</b>) ethylene glycol; (<b>b</b>) propylene glycol.</p>
Full article ">Figure 5
<p>(<b>a</b>) Thickness and (<b>b</b>) Electrical resistance of the silver films with respect to the number of laser scans; (<b>c</b>) EDS results for the silver lines with 10 scans.</p>
Full article ">Figure 6
<p>Silver lines fabricated on polyimide film substrate using LDSP (with ethylene glycol (EG) based solution): (<b>a</b>) microscopic picture of the silver line; (<b>b</b>) a demonstration of silver patterns on the flexible PI substrate.</p>
Full article ">Figure 7
<p>(<b>a</b>) Thickness of the silver film; (<b>b</b>) Reflectivity for 532 nm wavelength light <span class="html-italic">versus</span> number of laser scans (symbols: experimental data, lines: fitted curves).</p>
Full article ">Figure 8
<p>(<b>a</b>) The processing temperature for the first scan with respect to time along the scan path; (<b>b</b>) The simulated temperature profile on the polyimide surface.</p>
Full article ">Figure 9
<p>Variation of the silver line growth rate (in thickness) with respect to the processing temperature. The filled dots: the filled dots shown here were obtained using the fitted curve in <a href="#nanomaterials-06-00012-f007" class="html-fig">Figure 7</a>a.</p>
Full article ">Figure 10
<p>The velocity (arrows) and temperature (color surface) distribution near the laser focal spot, (<b>a</b>) Side view; (<b>b</b>) 3D view.</p>
Full article ">
2611 KiB  
Article
Repair of the Orbital Wall Fractures in Rabbit Animal Model Using Nanostructured Hydroxyapatite-Based Implant
by Sinziana Gradinaru, Laura Madalina Popescu, Roxana Mioara Piticescu, Sabina Zurac, Radu Ciuluvica, Alexandrina Burlacu, Raluca Tutuianu, Sorina-Nicoleta Valsan, Adrian Mihail Motoc and Liliana Mary Voinea
Nanomaterials 2016, 6(1), 11; https://doi.org/10.3390/nano6010011 - 7 Jan 2016
Cited by 9 | Viewed by 6130
Abstract
Cellular uptake and cytotoxicity of nanostructured hydroxyapatite (nanoHAp) are dependent on its physical parameters. Therefore, an understanding of both surface chemistry and morphology of nanoHAp is needed in order to be able to anticipate its in vivo behavior. The aim of this paper [...] Read more.
Cellular uptake and cytotoxicity of nanostructured hydroxyapatite (nanoHAp) are dependent on its physical parameters. Therefore, an understanding of both surface chemistry and morphology of nanoHAp is needed in order to be able to anticipate its in vivo behavior. The aim of this paper is to characterize an engineered nanoHAp in terms of physico-chemical properties, biocompatibility, and its capability to reconstitute the orbital wall fractures in rabbits. NanoHAp was synthesized using a high pressure hydrothermal method and characterized by physico-chemical, structural, morphological, and optical techniques. X-ray diffraction revealed HAp crystallites of 21 nm, while Scanning Electron Microscopy (SEM) images showed spherical shapes of HAp powder. Mean particle size of HAp measured by DLS technique was 146.3 nm. Biocompatibility was estimated by the effect of HAp powder on the adhesion and proliferation of mesenchymal stem cells (MSC) in culture. The results showed that cell proliferation on powder-coated slides was between 73.4% and 98.3% of control cells (cells grown in normal culture conditions). Computed tomography analysis of the preformed nanoHAp implanted in orbital wall fractures, performed at one and two months postoperative, demonstrated the integration of the implants in the bones. In conclusion, our engineered nanoHAp is stable, biocompatible, and may be safely considered for reconstruction of orbital wall fractures. Full article
(This article belongs to the Special Issue Nanoparticles Assisted Drug Delivery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) X-ray diffractogram of HAp nanopowder; (<b>b</b>) X-ray diffractogram of HAp sintered at 800 °C/30 min.</p>
Full article ">Figure 2
<p>(<b>a</b>) Scanning Electron Microscopy (SEM) image of HAp nanopowder; (<b>b</b>) SEM image of HAp sintered at 800 °C/30 min.</p>
Full article ">Figure 3
<p>Particle size distribution of HAp nanopowder.</p>
Full article ">Figure 4
<p>Mesenchymal stem cell (MSC) proliferation in the presence of HAp solubilized in phosphorous acid 5% (pH = 1.31, HAp1s) or phosphorous acid 1% (pH = 1.88, HAp2s). Results are given as mean ± standard deviation (S.D.) of the cell viability expressed as percentage of control cells (cells incubated in normal growth medium in the absence of HAp).</p>
Full article ">Figure 5
<p>3D computed tomography (CT) reconstruction image of the animal model before surgery. Scale was estimated based on corresponding 2-D image.</p>
Full article ">Figure 6
<p>3D CT osseous reconstruction images at one month postoperative for rabbits 1–4. Scale was estimated based on corresponding 2-D image.</p>
Full article ">Figure 7
<p>3D CT osseous reconstruction images at two months postoperative for rabbits 1–4. Scale was estimated based on corresponding 2-D image.</p>
Full article ">Figure 8
<p>(<b>a</b>) Microscopic aspects of HAp implant two months after implantation, as revealed by hematoxilin eosin staining (the magnification is 10×); (<b>b</b>) Detailed image of the inset in figure (a) at 20× magnification; (<b>c</b>) Detailed image of the inset in figure (b) at 40× magnification. Black arrowheads mark HAp implant. Black arrows mark the osseous tissue. White arrowheads mark osteoclasts. White arrows mark the inferior rectus muscle. Red arrowheads mark the adipose tissue adjacent to bone structure.</p>
Full article ">Figure 9
<p>CD31 immunohistochemical staining of the implant. Black arrowheads mark the vascular structures, black arrows mark the CD31-positive cells delineating the vascular structures. Red arrowheads mark the osteoclasts at the interface between the implant and bone. The magnification is 100×.</p>
Full article ">Figure 10
<p>Intra-operatory aspect: (<b>a</b>) transconjunctival approach and lateral cantothomy facilitate the inferior rim periosteal elevation; (<b>b</b>) the preformed osseous defect in the inferior orbital wall of 0.5 cm diameter (white arrow); (<b>c</b>) the orbital implant placed in the defect area and sutured to the periosteum; (<b>d</b>) the rest of the defect covered with hydroxyapatite powder.</p>
Full article ">
1729 KiB  
Article
The Influence of Carbonaceous Matrices and Electrocatalytic MnO2 Nanopowders on Lithium-Air Battery Performances
by Alessandro Minguzzi, Gianluca Longoni, Giuseppe Cappelletti, Eleonora Pargoletti, Chiara Di Bari, Cristina Locatelli, Marcello Marelli, Sandra Rondinini and Alberto Vertova
Nanomaterials 2016, 6(1), 10; https://doi.org/10.3390/nano6010010 - 6 Jan 2016
Cited by 17 | Viewed by 5813
Abstract
Here, we report new gas diffusion electrodes (GDEs) prepared by mixing two different pore size carbonaceous matrices and pure and silver-doped manganese dioxide nanopowders, used as electrode supports and electrocatalytic materials, respectively. MnO2 nanoparticles are finely characterized in terms of structural (X-ray [...] Read more.
Here, we report new gas diffusion electrodes (GDEs) prepared by mixing two different pore size carbonaceous matrices and pure and silver-doped manganese dioxide nanopowders, used as electrode supports and electrocatalytic materials, respectively. MnO2 nanoparticles are finely characterized in terms of structural (X-ray powder diffraction (XRPD), energy dispersive X-ray (EDX)), morphological (SEM, high-angle annular dark field (HAADF)-scanning transmission electron microscopy (STEM)/TEM), surface (Brunauer Emmet Teller (BET)-Barrett Joyner Halenda (BJH) method) and electrochemical properties. Two mesoporous carbons, showing diverse surface areas and pore volume distributions, have been employed. The GDE performances are evaluated by chronopotentiometric measurements to highlight the effects induced by the adopted materials. The best combination, hollow core mesoporous shell carbon (HCMSC) with 1.0% Ag-doped hydrothermal MnO2 (M_hydro_1.0%Ag) allows reaching very high specific capacity close to 1400 mAh·g−1. Considerably high charge retention through cycles is also observed, due to the presence of silver as a dopant for the electrocatalytic MnO2 nanoparticles. Full article
(This article belongs to the Special Issue Nanoparticles for Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>High-angle annular dark field (HAADF)-scanning transmission electron microscopy (STEM) (on the left) and transmission electron microscope (TEM) (on the right) images of (<b>a</b>,<b>b</b>) hydrothermal MnO<sub>2</sub> (M_hydro), (<b>c</b>,<b>d</b>) M_500 (500 °C) and (<b>e</b>,<b>f</b>) M_hydro_1.0%Ag samples.</p>
Full article ">Figure 2
<p>Comparison of the pore size distribution (calculated by the Barrett Joyner Halenda (BJH) method) of MnO<sub>2</sub> powders at increasing calcination temperature.</p>
Full article ">Figure 3
<p>X-ray powder diffraction (XRPD) pattern of the M_hydro sample with the most intense reflections (<span class="html-italic">hkl</span>, intensity) of the main polymorphs (γ-MnO<sub>2</sub> nsutite, β-MnO<sub>2</sub> ramsdellite (RAM), β-MnO<sub>2</sub> pyrolusite (PYR) and α-MnO<sub>2</sub> hollandite).</p>
Full article ">Figure 4
<p>XRPD patterns of MnO<sub>2</sub> samples calcined at different temperatures. The most significant reflections for γ-MnO<sub>2</sub> nsutite, β-MnO<sub>2</sub> pyrolusite, α-MnO<sub>2</sub> hollandite and Mn<sub>2</sub>O<sub>3</sub> bixbite are highlighted.</p>
Full article ">Figure 5
<p><span class="html-italic">E vs. Q</span> curves of GDEs by varying the (<b>a</b>) mesoporous carbon matrices and (<b>b</b>) carbonaceous supports added with MnO<sub>2</sub>, during the first discharge cycle carried out at 0.34 mA·cm<sup>−2</sup> in the Swagelok™-type cell (<span class="html-italic">S-</span>cell). Bold line: reference GDEs (GDE-V, Curve 1 and GDE-V-Pt, Curve 7).</p>
Full article ">Figure 6
<p>Specific capacity of GDE-H-M_hydro and GDE-H-M_hydro_1.0%Ag during the first five cycles in the home-made cell (<span class="html-italic">H</span>-cell) (black and grey histograms, respectively).</p>
Full article ">Figure 7
<p>Discharge and charge cycles for (<b>a</b>) GDE-H-M_hydro and (<b>b</b>) GDE-H-M_hydro_1.0%Ag in the <span class="html-italic">H</span>-cell.</p>
Full article ">Figure 8
<p>X-ray diffraction lines of (<b>a</b>) GDE-H-M_hydro and (<b>b</b>) GDE-H-M_hydro_1.0%Ag obtained by subtraction between the GDEs after five cycles and before cycling. The * label indicates the presence of MnO<sub>2</sub> polymorphs.</p>
Full article ">
3490 KiB  
Article
Aqueous Dispersions of Silica Stabilized with Oleic Acid Obtained by Green Chemistry
by Cristina Lavinia Nistor, Raluca Ianchis, Marius Ghiurea, Cristian-Andi Nicolae, Catalin-Ilie Spataru, Daniela Cristina Culita, Jeanina Pandele Cusu, Victor Fruth, Florin Oancea and Dan Donescu
Nanomaterials 2016, 6(1), 9; https://doi.org/10.3390/nano6010009 - 5 Jan 2016
Cited by 12 | Viewed by 7854
Abstract
The present study describes for the first time the synthesis of silica nanoparticles starting from sodium silicate and oleic acid (OLA). The interactions between OLA and sodium silicate require an optimal OLA/OLANa molar ratio able to generate vesicles that can stabilize silica particles [...] Read more.
The present study describes for the first time the synthesis of silica nanoparticles starting from sodium silicate and oleic acid (OLA). The interactions between OLA and sodium silicate require an optimal OLA/OLANa molar ratio able to generate vesicles that can stabilize silica particles obtained by the sol-gel process of sodium silicate. The optimal molar ratio of OLA/OLANa can be ensured by a proper selection of OLA and respectively of sodium silicate concentration. The titration of sodium silicate with OLA revealed a stabilization phenomenon of silica/OLA vesicles and the dependence between their average size and reagent’s molar ratio. Dynamic light scattering (DLS) and scanning electron microscopy (SEM) measurements emphasized the successful synthesis of silica nanoparticles starting from renewable materials, in mild condition of green chemistry. By grafting octadecyltrimethoxysilane on the initial silica particles, an increased interaction between silica particles and the OLA/OLANa complex was achieved. This interaction between the oleyl and octadecyl chains resulted in the formation of stable gel-like aqueous systems. Subsequently, olive oil and an oleophylic red dye were solubilized in these stable aqueous systems. This great dispersing capacity of oleosoluble compounds opens new perspectives for future green chemistry applications. After the removal of water and of the organic chains by thermal treatment, mesoporous silica was obtained. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>pH modification of the dispersions prepared at a fixed amount of sodium silicate and different quantities of oleic acid (OLA); <b>Inset Picture</b>: The dispersions prepared at different quantities of OLA and a fixed amount of sodium silicate (<b>1.</b> 0.125/1; <b>2.</b> 0.25/1; <b>3.</b> 0.5/1; <b>4.</b> 1/1; <b>5.</b> 1.5/1; <b>6.</b> 2/1; <b>7.</b> 3/1 and <b>8.</b> 4/1 OLA/Na).</p>
Full article ">Figure 2
<p>(<b>I</b>) Size modification of the dispersions prepared at a fixed amount of sodium silicate and different quantities of OLA; (<b>II</b>) Zeta potential modification of the dispersions prepared at a fixed amount of sodium silicate and different quantities of OLA (Dmed represents the medium/average diameter recorded for the prepared silica dispersions).</p>
Full article ">Figure 3
<p>Scanning electron microscopy (SEM) pictures of the samples obtained at (<b>I</b>) 1.5/1 (sample No.5) and (<b>II</b>) 3/1 OLA/Na (sample No.7).</p>
Full article ">Figure 4
<p>The modification of pH depending on the ODTMOS fraction; <b>Inset picture</b>: The dispersions prepared at OLA/Na 2/1 (mole/mole) and 0/1(<span class="html-italic">sample H</span>); 1/1 (<span class="html-italic">sample A</span>); 1/5 (<span class="html-italic">sample G</span>) and 1/10 (<span class="html-italic">sample E</span>) (ODTMOS/SiO<sub>2</sub> (mole/mole)).</p>
Full article ">Figure 5
<p>The modification of silica particles <b>(I</b>) average diameters and (<b>II</b>) zeta potential depending on the fraction of ODTMOS.</p>
Full article ">Figure 6
<p>SEM pictures of (<b>I</b>) <span class="html-italic">sample A</span> (1/1 ODTMOS/SiO<sub>2</sub>) and (<b>II</b>) <span class="html-italic">sample H</span> (0/1 ODTMOS/SiO<sub>2</sub>) at 2/1 OLA/Na.</p>
Full article ">Figure 7
<p>The melting enthalpy of 1 mole of hydrophobic component (OLA + ODTMOS) as function of molar ratio of OLA (<span class="html-italic">f</span><sub>OLA</sub> = mole OLA/mole (OLA + ODTMOS)).</p>
Full article ">Figure 8
<p>Differential scanning calorimetry (DSC) analysis of <span class="html-italic">sample A</span> (as synthesized and washed with ammonium hydroxide).</p>
Full article ">Figure 9
<p>(<b>a</b>) Nitrogen isotherms for calcined <span class="html-italic">samples</span> (<span class="html-italic">A</span>, <span class="html-italic">G</span> and <span class="html-italic">H</span>); (<b>b</b>) Barrett-Joyner-Halenda (BJH) desorption dV/dD pore volume for <span class="html-italic">A</span>, <span class="html-italic">G</span> and <span class="html-italic">H samples</span> (OLA/ Na = 2/1; ODTMOS /SiO<sub>2</sub> = 1/1, 1/5, 0/1).</p>
Full article ">Figure 10
<p>Schematic representation of the synthesis of mesoporous silica nanoparticles stabilized with an oleic acid/sodium-oleate complex starting from sodium silicate and octadecyltrimethoxysilane.</p>
Full article ">
4645 KiB  
Article
Targeted Sterically Stabilized Phospholipid siRNA Nanomedicine for Hepatic and Renal Fibrosis
by Fatima Khaja, Dulari Jayawardena, Antonina Kuzmis and Hayat Önyüksel
Nanomaterials 2016, 6(1), 8; https://doi.org/10.3390/nano6010008 - 5 Jan 2016
Cited by 20 | Viewed by 6565
Abstract
Since its discovery, small interfering RNA (siRNA) has been considered a potent tool for modulating gene expression. It has the ability to specifically target proteins via selective degradation of messenger RNA (mRNA) not easily accessed by conventional drugs. Hence, RNA interference (RNAi) therapeutics [...] Read more.
Since its discovery, small interfering RNA (siRNA) has been considered a potent tool for modulating gene expression. It has the ability to specifically target proteins via selective degradation of messenger RNA (mRNA) not easily accessed by conventional drugs. Hence, RNA interference (RNAi) therapeutics have great potential in the treatment of many diseases caused by faulty protein expression such as fibrosis and cancer. However, for clinical application siRNA faces a number of obstacles, such as poor in vivo stability, and off-target effects. Here we developed a unique targeted nanomedicine to tackle current siRNA delivery issues by formulating a biocompatible, biodegradable and relatively inexpensive nanocarrier of sterically stabilized phospholipid nanoparticles (SSLNPs). This nanocarrier is capable of incorporating siRNA in its core through self-association with a novel cationic lipid composed of naturally occuring phospholipids and amino acids. This overall assembly protects and delivers sufficient amounts of siRNA to knockdown over-expressed protein in target cells. The siRNA used in this study, targets connective tissue growth factor (CTGF), an important regulator of fibrosis in both hepatic and renal cells. Furthermore, asialoglycoprotein receptors are targeted by attaching the galactosamine ligand to the nanocarries which enhances the uptake of nanoparticles by hepatocytes and renal tubular epithelial cells, the major producers of CTGF in fibrosis. On animals this innovative nanoconstruct, small interfering RNA in sterically stabilized phospholipid nanoparticles (siRNA-SSLNP), showed favorable pharmacokinetic properties and accumulated mostly in hepatic and renal tissues making siRNA-SSLNP a suitable system for targeting liver and kidney fibrotic diseases. Full article
(This article belongs to the Special Issue Nanoparticles Assisted Drug Delivery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>small interfering RNA in sterically stabilized phospholipid nanoparticles (siRNA-SSLNP) optimization <span class="html-italic">in vitro</span>: Representative particle size distribution of (<b>A</b>) Unimodal size distribution of empty SSMM (sterically stabilized mixed micelles); (<b>B</b>) Empty SSMM and siRNA-SSLNP with N/P ratio of 10; (<b>C</b>) Empty SSMM and siRNA-SSLNP with N/P ratio of 20; (<b>D</b>) Empty SSMM and siRNA-SSLNP with N/P ratio of 30; (<b>E</b>) siRNA lipofectamine (siRNA-Lipofectamine) complex; (<b>F</b>) Free siRNA molecules.</p>
Full article ">Figure 2
<p>siRNA-SSLNP <span class="html-italic">in vitro</span> characterization: (<b>A</b>) Gel retardation assay of different formulations of siRNA, containing 200 nM siRNA per sample, on TBE-urea 15% gel , at a voltage of 180 V for 60 min, then stained with 1:500 SYBR Green-II in TBE with mild agitation for 30 min; (<b>B</b>) Fluorescence intensities measured by SYBR Green-II exclusion assay of SSMM and siRNA-SSLNP complexes at varying N/P ratios and siRNA with lipofectamine (LF) showing percent of un-incorporated siRNA (* <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs</span>. free siRNA; mean ± SD; <span class="html-italic">n</span> = 3 replicates/group); (<b>C</b>) Gel retardation assay of different siRNA formulations after treatment with RNase; (<b>D</b>) Fluorescence intensities measured by SYBR Green-II fluorescence assay of different siRNA formulations after treatment with RNase (* <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs</span>. free siRNA; † <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> siRNA-lipofectamine; mean ± SD; <span class="html-italic">n</span> = 3 replicates/group).</p>
Full article ">Figure 3
<p>Physicochemical characterization of siRNA-SSLNP-GalN: (<b>A</b>) Particle size distribution showing SSLNP-GalN peak at 91 ± 13 nm; (<b>B</b>) Transmission electron microscopy (TEM) image of siRNA-SSLNP-GalN, scale bar = 100 nm; (<b>C</b>) Results of SYBR Green-II exclusion assay. Bars represent percentage of siRNA before (un-incorporated) and after treatment with RNase enzyme (mean ± SD; <span class="html-italic">n</span> = 3 replicates/group).</p>
Full article ">Figure 4
<p>Cell uptake and cytotoxicity assays: (<b>A</b>) Hepatic Hep-G2 cell uptake of FAM-labeled siRNA in various complexes. Changes in FACS histogram indicative of siRNA positive cells (upper), bars represent quantitative analysis of FACS histogram as a percentage siRNA-positive cells (lower). (* <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> free siRNA and untreated control, † <span class="html-italic">p</span> &gt; 0.05 meaning no statistical significance <span class="html-italic">vs</span>. siRNA-LF treated cells); (<b>B</b>) Cytotoxicity of siRNA in various complexes against primary hepatic stellate cells HSC at different siRNA concentration as determined by membrane integrity (LDH) assay; (<b>C</b>) Relative Hep-G2 cell viability expressed as a percentage of untreated control as a measure of cytotoxicity of siRNA complexes using MTS assay after 72 h incubation; (<b>D</b>) Cytotoxicity of different siRNA formulations after incubation with renal HK-2 cells for 72 h; (<b>E</b>) Cell proliferation kinetics of HK-2 cells after treatment with different formulations at siRNA concentration equivalent to 250 nM assessed at 24, 48, and 72 h time points (* <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> siRNA-LF, data on B–D presented as mean ± SD; <span class="html-italic">n</span> = 3 replicates/group).</p>
Full article ">Figure 5
<p>Protein downregulation: (<b>A</b>) Reduction of connective tissue growth factor (CTGF) expression in human hepatic Hep-G2 cells 24 h post transfection with anti CTGF-siRNA in different complexes; (<b>B</b>) Reduction of extracellular matrix (ECM) collagen expression in primary human hepatic stellate cells (HSC) 24 h post transfection with anti CTGF-siRNA in different complexes; (<b>C</b>) GTGF protein downregulation in renal tubular HK-2 cells, activated with transforming growth factor β1 (TGF-β1), 72 h post transfection with anti CTGF-siRNA in different complexes; (<b>D</b>) Reduction in ECM collagen expression by TGFβ activated HK-2 cells 72 h post transfection with anti CTFG-siRNA in different complexes. (Data are expressed as percent of the untreated control; mean ± SD; <span class="html-italic">n</span> = 3/treatment; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs</span>. free siRNA at a corresponding siRNA dose; <sup>#</sup> <span class="html-italic">p</span> &gt; 0.05 or statistically not significant <span class="html-italic">vs.</span> siRNA-lipofectamine (LF); ns—non significant among the groups indicated). The scrambled siRNA treatment showed no significance as compared to untreated controls and therefore control refers to untreated controls (Data not plotted).</p>
Full article ">Figure 6
<p>Reversal of primary hepatic stellate cell (HSC) activation: (<b>A</b>) Down-regulation of collagen I; (<b>B</b>) Collagen III; and (<b>C</b>) α-smooth muscle actin (α-SMA) protein expression indicative of the reversal of activated myofibroblasts to quiescent stellate cells. Activated HSC were transfected with connective tissue growth factor (CTGF-siRNA) in various formulations. Standard immunocytochemistry performed 24 h post-transfection, cells were probed with primary antibodies followed by secondary Alexa-Fluor 488 (<b>green</b>) labeled antibody, then DAPI for nuclear staining (<b>blue</b>).</p>
Full article ">Figure 7
<p>Biodistribution of different siRNA formulations compared to free Cy5 fluorophore over 24 h periods in (<b>A</b>) liver; (<b>B</b>) lung; (<b>C</b>) spleen; (<b>D</b>) heart; and (<b>E</b>) kidneys. Targeted formulation (siRNA-SSLNP-GalN) shows significant concentrations in liver and kidneys over observation period (<span class="html-italic">n</span> = 4 for each time point; * <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> free siRNA treated animals, † <span class="html-italic">p</span> &lt; 0.05 <span class="html-italic">vs.</span> free Cy5 treated animals); (<b>F</b>) Plasma concentration <span class="html-italic">vs.</span> time after single intravenous administration of various Cy-5 labeled formulations in Balb/c mice. (Data are presented as mean ± SD; <span class="html-italic">n</span> = 4 animals/each time point, MFI-Mean fluorescence intensity).</p>
Full article ">
2558 KiB  
Article
Determination of Cd2+ and Pb2+ Based on Mesoporous Carbon Nitride/Self-Doped Polyaniline Nanofibers and Square Wave Anodic Stripping Voltammetry
by Chang Zhang, Yaoyu Zhou, Lin Tang, Guangming Zeng, Jiachao Zhang, Bo Peng, Xia Xie, Cui Lai, Beiqing Long and Jingjing Zhu
Nanomaterials 2016, 6(1), 7; https://doi.org/10.3390/nano6010007 - 4 Jan 2016
Cited by 39 | Viewed by 7621
Abstract
The fabrication and evaluation of a glassy carbon electrode (GCE) modified with self-doped polyaniline nanofibers (SPAN)/mesoporous carbon nitride (MCN) and bismuth for simultaneous determination of trace Cd2+ and Pb2+ by square wave anodic stripping voltammetry (SWASV) are presented here. The morphology [...] Read more.
The fabrication and evaluation of a glassy carbon electrode (GCE) modified with self-doped polyaniline nanofibers (SPAN)/mesoporous carbon nitride (MCN) and bismuth for simultaneous determination of trace Cd2+ and Pb2+ by square wave anodic stripping voltammetry (SWASV) are presented here. The morphology properties of SPAN and MCN were characterized by transmission electron microscopy (TEM), and the electrochemical properties of the fabricated electrode were characterized by cyclic voltammetry (CV). Experimental parameters, such as deposition time, pulse potential, step potential, bismuth concentration and NaCl concentration, were optimized. Under the optimum conditions, the fabricated electrode exhibited linear calibration curves ranging from 5 to 80 nM for Cd2+ and Pb2+. The limits of detection (LOD) were 0.7 nM for Cd2+ and 0.2 nM for Pb2+ (S/N = 3). Additionally, the repeatability, reproducibility, anti-interference ability and application were also investigated, and the proposed electrode exhibited excellent performance. The proposed method could be extended for other heavy metal determination. Full article
(This article belongs to the Special Issue Nanomaterials for Biosensing Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Transmission electron microscopy (TEM) of the (<b>A</b>) self-doped polyaniline (SPAN) nanofiber and (<b>B</b>) mesoporous carbon nitride (MCN).</p>
Full article ">Figure 2
<p>(<b>A</b>) Cyclic voltammetry diagrams of the glassy carbon electrode (GCE), GCE/SPAN and GCE/SPAN/MCN, using a 0.1 M KCl solution containing 5.0 mM ferro-/ferri-cyanide, with a potential range of −0.4–0.8 V and a scan rate of 100 mV·s<sup>−1</sup>; (<b>B</b>) electrochemical impedance spectra of GCE, GCE/SPAN and GCE/SPAN/MCN using a 0.1 M KCl solution containing 5.0 mM ferro-/ferri-cyanide, with a frequency range of 0.1–10<sup>5</sup> Hz, a bias potential of 0.19 V <span class="html-italic">vs.</span> a saturated calomel electrode (SCE) and an alternating current (AC) amplitude of 5 mV.</p>
Full article ">Figure 3
<p>(<b>A</b>) Bi<sup>3+</sup> concentration effect on peak height in a solution containing 200 mM NaCl, 10 μg·L<sup>−1</sup> Pb<sup>2+</sup> and Cd<sup>2+</sup>. Square wave anodic stripping voltammetry (SWASV) parameters: <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 50 Hz, deposition time = 300 s and equilibrium time = 20 s. (<b>B</b>) NaCl concentration effect on peak height in a solution containing 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, 10 μg·L<sup>−1</sup> Pb<sup>2+</sup> and Cd<sup>2+</sup>. SWASV parameters: <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 50 Hz, deposition time = 300 s and equilibrium time = 20 s.</p>
Full article ">Figure 4
<p>(<b>A</b>) Frequency effect on peak height in a solution containing 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, 10 μg·L<sup>−1</sup> Pb<sup>2+</sup> and Cd<sup>2+</sup> and 300 mM NaCl. SWASV parameters: <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 1–100 Hz, deposition time = 600 s and equilibrium time = 20 s. (<b>B</b>) Optimization of pulse potential (<span class="html-italic">E</span><sub>pulse</sub>) using a solution containing 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, 10 μg·L<sup>−1</sup> Pb<sup>2+</sup> and Cd<sup>2+</sup> and 300 mM NaCl. SWV parameters: <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.005–0.05 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 100 Hz, deposition time = 600 s and equilibrium time = 20 s.</p>
Full article ">Figure 5
<p>(<b>A</b>) SWASV curves at GCE/SPAN/MCN in 10 mM acetate buffer (pH 4.6) containing 300 mM NaCl, 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, with different Cd<sup>2+</sup> and Pb<sup>2+</sup> concentrations (from a–i: 0, 1, 2, 5, 10, 20 , 40, 60, 80 μg·L<sup>−1</sup>). The SWASV parameters were <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 100 Hz, deposition time = 300 s and equilibrium time = 20 s. (<b>B</b>) and (<b>C</b>) show the plots of stripping peak current <span class="html-italic">vs.</span> the Cd(II) and Pb(II) concentration.</p>
Full article ">Figure 5 Cont.
<p>(<b>A</b>) SWASV curves at GCE/SPAN/MCN in 10 mM acetate buffer (pH 4.6) containing 300 mM NaCl, 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, with different Cd<sup>2+</sup> and Pb<sup>2+</sup> concentrations (from a–i: 0, 1, 2, 5, 10, 20 , 40, 60, 80 μg·L<sup>−1</sup>). The SWASV parameters were <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 100 Hz, deposition time = 300 s and equilibrium time = 20 s. (<b>B</b>) and (<b>C</b>) show the plots of stripping peak current <span class="html-italic">vs.</span> the Cd(II) and Pb(II) concentration.</p>
Full article ">Figure 5 Cont.
<p>(<b>A</b>) SWASV curves at GCE/SPAN/MCN in 10 mM acetate buffer (pH 4.6) containing 300 mM NaCl, 300 μg·L<sup>−1</sup> Bi<sup>3+</sup>, with different Cd<sup>2+</sup> and Pb<sup>2+</sup> concentrations (from a–i: 0, 1, 2, 5, 10, 20 , 40, 60, 80 μg·L<sup>−1</sup>). The SWASV parameters were <span class="html-italic">E</span><sub>begin</sub> = −1 V, <span class="html-italic">E</span><sub>end</sub> = −0.4 V, <span class="html-italic">E</span><sub>step</sub> = 0.010 V, <span class="html-italic">E</span><sub>pulse</sub> = 0.02 V, <span class="html-italic">E</span><sub>condition</sub> = −0.4 V, <span class="html-italic">E</span><sub>deposition</sub> = −1 V, frequency = 100 Hz, deposition time = 300 s and equilibrium time = 20 s. (<b>B</b>) and (<b>C</b>) show the plots of stripping peak current <span class="html-italic">vs.</span> the Cd(II) and Pb(II) concentration.</p>
Full article ">
6556 KiB  
Article
Effects of Thickness and Amount of Carbon Nanofiber Coated Carbon Fiber on Improving the Mechanical Properties of Nanocomposites
by Ferial Ghaemi, Ali Ahmadian, Robiah Yunus, Fudziah Ismail and Saeed Rahmanian
Nanomaterials 2016, 6(1), 6; https://doi.org/10.3390/nano6010006 - 2 Jan 2016
Cited by 18 | Viewed by 5546
Abstract
In the current study, carbon nanofibers (CNFs) were grown on a carbon fiber (CF) surface by using the chemical vapor deposition method (CVD) and the influences of some parameters of the CVD method on improving the mechanical properties of a polypropylene (PP) composite [...] Read more.
In the current study, carbon nanofibers (CNFs) were grown on a carbon fiber (CF) surface by using the chemical vapor deposition method (CVD) and the influences of some parameters of the CVD method on improving the mechanical properties of a polypropylene (PP) composite were investigated. To obtain an optimum surface area, thickness, and yield of the CNFs, the parameters of the chemical vapor deposition (CVD) method, such as catalyst concentration, reaction temperature, reaction time, and hydrocarbon flow rate, were optimized. It was observed that the optimal surface area, thickness, and yield of the CNFs caused more adhesion of the fibers with the PP matrix, which enhanced the composite properties. Besides this, the effectiveness of reinforcement of fillers was fitted with a mathematical model obtaining good agreement between the experimental result and the theoretical prediction. By applying scanning electronic microscope (SEM), transmission electron microscope (TEM), and Raman spectroscopy, the surface morphology and structural information of the resultant CF-CNF were analyzed. Additionally, SEM images and a mechanical test of the composite with a proper layer of CNFs on the CF revealed not only a compactness effect but also the thickness and surface area roles of the CNF layers in improving the mechanical properties of the composites. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scanning electron microscopy (SEM) image; (<b>b</b>) Raman spectroscopy of pristine carbon fiber.</p>
Full article ">Figure 2
<p>SEM images and Raman spectra of different agglomerations of grown carbon nanofibers (CNFs) on carbon fiber (CF) at (<b>a</b>) 50 mM, (<b>b</b>) 100 mM, and (<b>c</b>) 150 mM catalyst concentrations at 550 °C for 30 min run time under 50 sccm acetylene flow rate.</p>
Full article ">Figure 3
<p>SEM images and Raman spectra of grown CNF by use of 100 mM catalyst concentration for 30 min under 50 sccm acetylene flow rate at (<b>a</b>) 450 °C, (<b>b</b>) 550 °C, and (<b>c</b>) 650 °C.</p>
Full article ">Figure 4
<p>Themogravimetric analysis (TGA) analysis of (<b>a</b>) neat CF and CF-CNF at (<b>b</b>) 450 °C, (<b>c</b>) 550 °C, and (<b>d</b>) 650 °C.</p>
Full article ">Figure 5
<p>SEM images and Raman Spectrum of CNF morphologies at (<b>a</b>) 10 min, (<b>b</b>) 30 min, and (<b>c</b>) 50 min using 100 mM acid concentration at 550 °C under 50 sccm acetylene flow rate.</p>
Full article ">Figure 6
<p>SEM images and Raman spectroscopy of carbon nanofiber on CF by use of 100 mM catalyst concentration at 550 °C for 30 min at (<b>a</b>) 25 sccm, (<b>b</b>) 50 sccm, and (<b>c</b>) 100 sccm flow rate of C<sub>2</sub>H<sub>2</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) SEM and (<b>b</b>) transmission electron microscopy (TEM) micrographs of optimum CNF.</p>
Full article ">Figure 8
<p>Effective reinforcement modulus of different fillers in polypropylene matrix (dark purple states minimum amount and light purple reveals maximum amount).</p>
Full article ">Figure 9
<p>SEM micrographs of fractured surface of (<b>a</b>) CF-CNF<sub>L</sub>/PP, (<b>b</b>) CF-CNF<sub>M</sub>/PP, and (<b>c</b>) CF-CNF<sub>H</sub>/PP composites.</p>
Full article ">
968 KiB  
Review
Biosensors Incorporating Bimetallic Nanoparticles
by John Rick, Meng-Che Tsai and Bing Joe Hwang
Nanomaterials 2016, 6(1), 5; https://doi.org/10.3390/nano6010005 - 31 Dec 2015
Cited by 56 | Viewed by 8797
Abstract
This article presents a review of electrochemical bio-sensing for target analytes based on the use of electrocatalytic bimetallic nanoparticles (NPs), which can improve both the sensitivity and selectivity of biosensors. The review moves quickly from an introduction to the field of bio-sensing, to [...] Read more.
This article presents a review of electrochemical bio-sensing for target analytes based on the use of electrocatalytic bimetallic nanoparticles (NPs), which can improve both the sensitivity and selectivity of biosensors. The review moves quickly from an introduction to the field of bio-sensing, to the importance of biosensors in today’s society, the nature of the electrochemical methods employed and the attendant problems encountered. The role of electrocatalysts is introduced with reference to the three generations of biosensors. The contributions made by previous workers using bimetallic constructs, grouped by target analyte, are then examined in detail; following which, the synthesis and characterization of the catalytic particles is examined prior to a summary of the current state of endeavor. Finally, some perspectives for the future of bimetallic NPs in biosensors are given. Full article
(This article belongs to the Special Issue Nanomaterials for Biosensing Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Generalized biosensor schematic showing stages in response to analyte.</p>
Full article ">Figure 2
<p>Showing (<b>a</b>) amperometric and (<b>b</b>) potentiometric electrode configurations.</p>
Full article ">Figure 3
<p>The three generations of biosensors showing transition from: (<b>a</b>) enzyme dependency to (<b>b</b>) mediator use and (<b>c</b>) finally to the use of a catalytic electrode.</p>
Full article ">Figure 4
<p>Catalytic structures: (<b>a</b>) heterogeneous, (<b>b</b>) core-shelled and (<b>c</b>) alloyed.</p>
Full article ">
3433 KiB  
Article
Dense Plasma Focus-Based Nanofabrication of III–V Semiconductors: Unique Features and Recent Advances
by Onkar Mangla, Savita Roy and Kostya (Ken) Ostrikov
Nanomaterials 2016, 6(1), 4; https://doi.org/10.3390/nano6010004 - 29 Dec 2015
Cited by 18 | Viewed by 7017
Abstract
The hot and dense plasma formed in modified dense plasma focus (DPF) device has been used worldwide for the nanofabrication of several materials. In this paper, we summarize the fabrication of III–V semiconductor nanostructures using the high fluence material ions produced by hot, [...] Read more.
The hot and dense plasma formed in modified dense plasma focus (DPF) device has been used worldwide for the nanofabrication of several materials. In this paper, we summarize the fabrication of III–V semiconductor nanostructures using the high fluence material ions produced by hot, dense and extremely non-equilibrium plasma generated in a modified DPF device. In addition, we present the recent results on the fabrication of porous nano-gallium arsenide (GaAs). The details of morphological, structural and optical properties of the fabricated nano-GaAs are provided. The effect of rapid thermal annealing on the above properties of porous nano-GaAs is studied. The study reveals that it is possible to tailor the size of pores with annealing temperature. The optical properties of these porous nano-GaAs also confirm the possibility to tailor the pore sizes upon annealing. Possible applications of the fabricated and subsequently annealed porous nano-GaAs in transmission-type photo-cathodes and visible optoelectronic devices are discussed. These results suggest that the modified DPF is an effective tool for nanofabrication of continuous and porous III–V semiconductor nanomaterials. Further opportunities for using the modified DPF device for the fabrication of novel nanostructures are discussed as well. Full article
(This article belongs to the Special Issue Plasma Nanoengineering and Nanofabrication)
Show Figures

Figure 1

Figure 1
<p>Atomic force microscopy (AFM) images of gallium arsenide (GaAs) nanostructures for (<b>a</b>) one, (<b>b</b>) two, (<b>c</b>) three shots on a silicon substrate placed at a distance of 4.0 cm.</p>
Full article ">Figure 2
<p>Typical photoluminescence (PL) spectrum of GaAs nanostructures deposited on a silicon substrate placed at a distance of 4.0 cm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Absorption spectra of GaAs nanostructures deposited on silicon substrate placed at a distance of 4.0 cm and 5.0 cm. Tauc plot showing band gap for (<b>b</b>) 4.0 cm, (<b>c</b>) 5.0 cm distances.</p>
Full article ">Figure 4
<p>Scanning electron microscopy (SEM) images of as-deposited porous GaAs (sample A). (<b>a</b>) surface, (<b>b</b>) cross-section view, (<b>c</b>) cross-section view showing upper and lower portions, (<b>d</b>) pores shown by arrow on surface.</p>
Full article ">Figure 5
<p>SEM images of 100 °C annealed porous GaAs (sample B). (<b>a</b>) surface, (<b>b</b>) cross-section view.</p>
Full article ">Figure 6
<p>SEM images of 200 °C annealed porous GaAs (sample C). (<b>a</b>) surface, (<b>b</b>) cross-section view.</p>
Full article ">Figure 7
<p>X-ray diffraction (XRD) pattern of as-deposited and rapid thermal annealing (RTA) porous GaAs.</p>
Full article ">Figure 8
<p>PL spectra of as-deposited and RTA porous GaAs.</p>
Full article ">Figure 9
<p>Transmission spectra of as-deposited and RTA porous GaAs.</p>
Full article ">Figure 10
<p>Schematic of modified dense plasma focus device.</p>
Full article ">
889 KiB  
Review
Receptor-Mediated Drug Delivery Systems Targeting to Glioma
by Shanshan Wang, Ying Meng, Chengyi Li, Min Qian and Rongqin Huang
Nanomaterials 2016, 6(1), 3; https://doi.org/10.3390/nano6010003 - 28 Dec 2015
Cited by 72 | Viewed by 11179
Abstract
Glioma has been considered to be the most frequent primary tumor within the central nervous system (CNS). The complexity of glioma, especially the existence of the blood-brain barrier (BBB), makes the survival and prognosis of glioma remain poor even after a standard treatment [...] Read more.
Glioma has been considered to be the most frequent primary tumor within the central nervous system (CNS). The complexity of glioma, especially the existence of the blood-brain barrier (BBB), makes the survival and prognosis of glioma remain poor even after a standard treatment based on surgery, radiotherapy, and chemotherapy. This provides a rationale for the development of some novel therapeutic strategies. Among them, receptor-mediated drug delivery is a specific pattern taking advantage of differential expression of receptors between tumors and normal tissues. The strategy can actively transport drugs, such as small molecular drugs, gene medicines, and therapeutic proteins to glioma while minimizing adverse reactions. This review will summarize recent progress on receptor-mediated drug delivery systems targeting to glioma, and conclude the challenges and prospects of receptor-mediated glioma-targeted therapy for future applications. Full article
(This article belongs to the Special Issue Nanoparticles in Theranostics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of single ligand-modified targeting systems.</p>
Full article ">Figure 2
<p>Design of the DOX-loaded GSPI-based system as a multifunctional drug delivery system for combined chemo-photothermal targeted glioma therapy.</p>
Full article ">
4545 KiB  
Article
Effects of Particle Hydrophobicity, Surface Charge, Media pH Value and Complexation with Human Serum Albumin on Drug Release Behavior of Mitoxantrone-Loaded Pullulan Nanoparticles
by Xiaojun Tao, Shu Jin, Dehong Wu, Kai Ling, Liming Yuan, Pingfa Lin, Yongchao Xie and Xiaoping Yang
Nanomaterials 2016, 6(1), 2; https://doi.org/10.3390/nano6010002 - 25 Dec 2015
Cited by 34 | Viewed by 7896
Abstract
We prepared two types of cholesterol hydrophobically modified pullulan nanoparticles (CHP) and carboxyethyl hydrophobically modified pullulan nanoparticles (CHCP) substituted with various degrees of cholesterol, including 3.11, 6.03, 6.91 and 3.46 per polymer, and named CHP−3.11, CHP−6.03, CHP−6.91 and [...] Read more.
We prepared two types of cholesterol hydrophobically modified pullulan nanoparticles (CHP) and carboxyethyl hydrophobically modified pullulan nanoparticles (CHCP) substituted with various degrees of cholesterol, including 3.11, 6.03, 6.91 and 3.46 per polymer, and named CHP−3.11, CHP−6.03, CHP−6.91 and CHCP−3.46. Dynamic laser light scattering (DLS) showed that the pullulan nanoparticles were 80–120 nm depending on the degree of cholesterol substitution. The mean size of CHCP nanoparticles was about 160 nm, with zeta potential −19.9 mV, larger than CHP because of the carboxyethyl group. A greater degree of cholesterol substitution conferred greater nanoparticle hydrophobicity. Drug-loading efficiency depended on nanoparticle hydrophobicity, that is, nanoparticles with the greatest degree of cholesterol substitution (6.91) showed the most drug encapsulation efficiency (90.2%). The amount of drug loading increased and that of drug release decreased with enhanced nanoparticle hydrophobicity. Nanoparticle surface-negative charge disturbed the amount of drug loading and drug release, for an opposite effect relative to nanoparticle hydrophobicity. The drug release in pullulan nanoparticles was higher pH 4.0 than pH 6.8 media. However, the changed drug release amount was not larger for negative-surface nanoparticles than CHP nanoparticles in the acid release media. Drug release of pullulan nanoparticles was further slowed with human serum albumin complexation and was little affected by nanoparticle hydrophobicity and surface negative charge. Full article
(This article belongs to the Special Issue Nanoparticles Assisted Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of pullulan.</p>
Full article ">Figure 2
<p>Structural graph of pullulan nanoparticle.</p>
Full article ">Figure 3
<p>Sketch illustration of human serum albumin (HSA) complexation with pullulan nanoparticles.</p>
Full article ">Figure 4
<p>The optical appearance of pullulan nanoparticles with different hydrophobicity and surface charge.</p>
Full article ">Figure 5
<p>Transmission electron microscopy (TEM) of CHP<sub>−6.91</sub> (<b>a</b>); CHP<sub>−6.03</sub> (<b>b</b>); CHP<sub>−3.11</sub> (<b>c</b>) and CHCP (<b>d</b>) nanoparticles.</p>
Full article ">Figure 6
<p>Size distribution of CHP<sub>−6.91</sub> (<b>a</b>); CHP<sub>−6.03</sub> (<b>b</b>); CHP<sub>−3.11</sub> (<b>c</b>) and CHCP (<b>d</b>) nanoparticles.</p>
Full article ">Figure 7
<p>Zeta potential of CHP<sub>−6.91</sub> (<b>a</b>); CHP<sub>−6.03</sub> (<b>b</b>); CHP<sub>−3.11</sub> (<b>c</b>) and CHCP (<b>d</b>) nanoparticles.</p>
Full article ">Figure 8
<p>TEM of CHP<sub>−6.03</sub> nanoparticles loading mitoxantrone.</p>
Full article ">Figure 9
<p>The mitoxantrone (MTO) release of pullulan nanoparticles in phosphate buffered saline (PBS) at 37 °C <span class="html-italic">in vitro</span> (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="black"> <mo>■</mo> </mstyle> </mrow> </semantics> </math>: free mitoxantrone, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#00B050"> <mo>▼</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−3.11</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#7030A0"> <mo>◄</mo> </mstyle> </mrow> </semantics> </math>: CHCP, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−6.03</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#0070C0"> <mo>▲</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−6.91</sub>).</p>
Full article ">Figure 10
<p>The mitoxantrone (MTO) release from pullulan nanoparticles in PBS buffer (pH 6.8) at 37 °C <span class="html-italic">in vitro</span> and acetate buffer (pH 4.0) (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#3333FF"> <mo>▲</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−3.11</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#00B0F0"> <mo>▼</mo> </mstyle> </mrow> </semantics> </math>: CHCP; pH 6.8), (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="black"> <mo>■</mo> </mstyle> </mrow> </semantics> </math>: CHCP, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−3.11</sub>; pH 4.0).</p>
Full article ">Figure 11
<p>The mitoxantrone (MTO) release of pullulan nanoparticles upon human serum albumin (HSA) complexation in PBS at 37 °C <span class="html-italic">in vitro</span> (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="red"> <mo>●</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−6.91</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#0070C0"> <mo>▲</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−3.11</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#00B050"> <mo>▼</mo> </mstyle> </mrow> </semantics> </math>: CHCP); and in the HSA release media (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="black"> <mo>■</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−6.91</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="#76923C"> <mo>►</mo> </mstyle> </mrow> </semantics> </math>: CHP<sub>−3.11</sub>, <math display="inline"> <semantics> <mrow> <mstyle mathcolor="fuchsia"> <mo>◄</mo> </mstyle> </mrow> </semantics> </math>: CHCP).</p>
Full article ">Figure 12
<p>Drug released from nanoparticles upon HSA complexation and in the HSA release media.</p>
Full article ">
4446 KiB  
Article
Synthesis and Application of Amine Functionalized Iron Oxide Nanoparticles on Menaquinone-7 Fermentation: A Step towards Process Intensification
by Alireza Ebrahiminezhad, Vikas Varma, Shuyi Yang, Younes Ghasemi and Aydin Berenjian
Nanomaterials 2016, 6(1), 1; https://doi.org/10.3390/nano6010001 - 25 Dec 2015
Cited by 60 | Viewed by 7256
Abstract
Industrial production of menaquione-7 by Bacillus subtilis natto is associated with major drawbacks. To address the current challenges in menaquione-7 fermentation, studying the effect of magnetic nanoparticles on the bacterial cells can open up a new domain for intensified menqainone-7 process. This article [...] Read more.
Industrial production of menaquione-7 by Bacillus subtilis natto is associated with major drawbacks. To address the current challenges in menaquione-7 fermentation, studying the effect of magnetic nanoparticles on the bacterial cells can open up a new domain for intensified menqainone-7 process. This article introduces the new concept of production and application of l-lysine coated iron oxide nanoparticles (l-Lys@IONs) as a novel tool for menaquinone-7 biosynthesis. l-Lys@IONs with the average size of 7 nm were successfully fabricated and were examined in a fermentation process of l-Lys@IONs decorated Bacillus subtilis natto. Based on the results, higher menaquinone-7 specific yield was observed for l-Lys@IONs decorated bacterial cells as compared to untreated bacteria. In addition, more than 92% removal efficacy was achieved by using integrated magnetic separation process. The present study demonstrates that l-Lys@IONs can be successfully applied during a fermentation of menaquinone-7 without any negative consequences on the culture conditions. This study provides a novel biotechnological application for IONs and their future role in bioprocess intensification. Full article
Show Figures

Figure 1

Figure 1
<p>Transmission electron micrographs of <span class="html-small-caps">l</span>-lysine coated magnetite nanoparticles.</p>
Full article ">Figure 2
<p>Fourier transform infrared spectroscopy (FTIR) spectra of (<b>a</b>) <span class="html-small-caps">l</span>-lysine coated magnetite nanoparticles and (<b>b</b>) pure <span class="html-small-caps">l</span>-lysine.</p>
Full article ">Figure 3
<p>Differential scanning calorimetry (DSC) curves of <span class="html-small-caps">l</span>-lysine coated magnetite nanoparticles.</p>
Full article ">Figure 4
<p>Vibrating sample magnetometer (VSM) diagrams of <span class="html-small-caps">l</span>-lysine coated magnetite nanoparticles.</p>
Full article ">Figure 5
<p>X-ray power diffraction patterns of <span class="html-small-caps">l</span>-lysine coated magnetite nanoparticles.</p>
Full article ">Figure 6
<p>SEM image of the produced <span class="html-italic">Bacillus subtilis natto</span> cells (<b>a</b>) untreated and (<b>b</b>) decorated with <span class="html-small-caps">l</span>-lysine-IONs.</p>
Full article ">Figure 7
<p><span class="html-italic">Bacillus subtilis natto</span> cells growth at different at <span class="html-small-caps">l</span>-lysine-IONs concentrations.</p>
Full article ">Figure 8
<p>MK-7 production at different at <span class="html-small-caps">l</span>-lysine-IONs concentrations.</p>
Full article ">Figure 9
<p>MK-7 specific yield at different at <span class="html-small-caps">l</span>-lysine-IONs concentrations.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop