[go: up one dir, main page]

Next Issue
Volume 6, September
Previous Issue
Volume 6, July
 
 

Nanomaterials, Volume 6, Issue 8 (August 2016) – 17 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:
7675 KiB  
Article
Study on Utilization of Carboxyl Group Decorated Carbon Nanotubes and Carbonation Reaction for Improving Strengths and Microstructures of Cement Paste
by Xiantong Yan, Hongzhi Cui, Qinghua Qin, Waiching Tang and Xiangming Zhou
Nanomaterials 2016, 6(8), 153; https://doi.org/10.3390/nano6080153 - 19 Aug 2016
Cited by 26 | Viewed by 6130 | Correction
Abstract
Carbon nanotubes (CNTs) have excellent mechanical properties and can be used to reinforce cement-based materials. On the other hand, the reaction product of carbonation with hydroxides in hydrated cement paste can reduce the porosity of cement-based materials. In this study, a novel method [...] Read more.
Carbon nanotubes (CNTs) have excellent mechanical properties and can be used to reinforce cement-based materials. On the other hand, the reaction product of carbonation with hydroxides in hydrated cement paste can reduce the porosity of cement-based materials. In this study, a novel method to improve the strength of cement paste was developed through a synergy of carbon nanotubes decorated with carboxyl group and carbonation reactions. The experimental results showed that the carboxyl group (–COOH) of decorated carbon nanotubes and the surfactant can control the morphology of the calcium carbonate crystal of carbonation products in hydrated cement paste. The spindle-like calcium carbonate crystals showed great morphological differences from those observed in the conventional carbonation of cement paste. The spindle-like calcium carbonate crystals can serve as fiber-like reinforcements to reinforce the cement paste. By the synergy of the carbon nanotubes and carbonation reactions, the compressive and flexural strengths of cement paste were significantly improved and increased by 14% and 55%, respectively, when compared to those of plain cement paste. Full article
(This article belongs to the Special Issue Nanomechanics of Carbon Nanotubes and Graphene Sheets)
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscope (SEM) images of CNT-CP microstructures: (<b>a</b>) Microstructure of crack-bridging CNTs in CNT-CP unbroken sample; (<b>b</b>) A close-up image of the microstructure of the image in <a href="#nanomaterials-06-00153-f001" class="html-fig">Figure 1</a>a displaying the bond between the CNT and C-S-H gel (main product of cement hydration).</p>
Full article ">Figure 2
<p>(<b>a</b>) Crack-bridging function of CNTs in a broken CNT-CP sample; (<b>b</b>) Pulled-out CNTs in a broken CNT-CP sample and the cement hydration product coated with CNT bundles.</p>
Full article ">Figure 3
<p>SEM image of the C-CP unbroken sample.</p>
Full article ">Figure 4
<p>SEM image of C-CNT-CP unbroken sample.</p>
Full article ">Figure 5
<p>Crystal structures of the polymorphs of calcium carbonate. (<b>a</b>) Calcite; (<b>b</b>) Aragonite; (<b>c</b>) Vaterite.</p>
Full article ">Figure 6
<p>SEM image of C-CNT-CP broken sample.</p>
Full article ">Figure 7
<p>Undispersed carbon nanotubes (CNTs).</p>
Full article ">
8681 KiB  
Article
A Novel Polyvinylidene Fluoride Tree-Like Nanofiber Membrane for Microfiltration
by Zongjie Li, Weimin Kang, Huihui Zhao, Min Hu, Na Wei, Jiuan Qiu and Bowen Cheng
Nanomaterials 2016, 6(8), 152; https://doi.org/10.3390/nano6080152 - 19 Aug 2016
Cited by 33 | Viewed by 5865
Abstract
A novel polyvinylidene fluoride (PVDF) tree-like nanofiber membrane (PVDF-TLNM) was fabricated by adding tetrabutylammonium chloride (TBAC) into a PVDF spinning solution via one-step electrospinning. The structure of the prepared membranes was characterized by field emission scanning electron microscopy (FE-SEM), Fourier transform infrared spectroscopy [...] Read more.
A novel polyvinylidene fluoride (PVDF) tree-like nanofiber membrane (PVDF-TLNM) was fabricated by adding tetrabutylammonium chloride (TBAC) into a PVDF spinning solution via one-step electrospinning. The structure of the prepared membranes was characterized by field emission scanning electron microscopy (FE-SEM), Fourier transform infrared spectroscopy (FT-IR) and pore size analysis, and the hydrophilic property and microfiltration performance were also evaluated. The results showed that the tree-like nanofiber was composed of trunk fibers and branch fibers with diameters of 100–500 nm and 5–100 nm, respectively. The pore size of PVDF-TLNM (0.36 μm) was smaller than that of a common nanofiber membrane (3.52 μm), and the hydrophilic properties of the membranes were improved significantly. The PVDF-TLNM with a thickness of 30 ± 2 μm showed a satisfactory retention ratio of 99.9% against 0.3 μm polystyrene (PS) particles and a high pure water flux of 2.88 × 104 L·m−2·h−1 under the pressure of 25 psi. This study highlights the potential benefits of this novel PVDF tree-like nanofiber membrane in the membrane field, which can achieve high flux rates at low pressure. Full article
(This article belongs to the Special Issue Environmental Applications and Implications of Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Field emission scanning electron microscopy (FE-SEM) images of polyvinylidene fluoride (PVDF) nanofiber membranes with different tetrabutylammonium chloride (TBAC) concentration: (<b>a</b>) no salt (pure PVDF); (<b>b</b>) 0.05 mol·L<sup>−1</sup> polyvinylidene fluoride tree-like nanofiber membrane-1 (PVDF-TLNMs-1); (<b>c</b>) 0.10 mol·L<sup>−1</sup> (PVDF-TLNMs-2); and (<b>d</b>) 0.15 mol·L<sup>−1</sup> (PVDF-TLNMs-3) (the inset is the pore size distribution of the membrane).</p>
Full article ">Figure 2
<p>Schematic diagram for the electrospinning process of (<b>a</b>) pure PVDF solution and (<b>b</b>) PVDF/TBAC solution; (<b>c</b>) FE-SEM images of polyvinylidene fluoride nanofiber membranes (PVDF-NMs) and (<b>d</b>) PVDF-TLNMs.</p>
Full article ">Figure 3
<p>Fourier transform infrared spectroscopy (FT-IR) spectroscopy for: (<b>a</b>) PVDF-NMs; (<b>b</b>) PVDF-TLNMs-3.</p>
Full article ">Figure 4
<p>Contact angle of PVDF nanofiber membranes with different contents of TBAC.</p>
Full article ">Figure 5
<p>FE-SEM images of the top surface, bottom surface and cross-section of the membranes after a filtration test: (<b>a1</b>–<b>a3</b>) PVDF-NMs; (<b>b1</b>–<b>b3</b>) PVDF-TLNMs-1; (<b>c1</b>–<b>c3</b>) PVDF-TLNMs-2; (<b>d1</b>–<b>d3</b>) PVDF-TLNMs-3 (inset: photographs of the filtrate solutions).</p>
Full article ">Figure 6
<p>The pure water flux of PVDF-NMs, PVDF-TLNMs-1 (0.05 mol·L<sup>−1</sup>), PVDF-TLNMs-2 (0.10 mol·L<sup>−1</sup>) and PVDF-TLNMs-3 (0.15 mol·L<sup>−1</sup>).</p>
Full article ">
1452 KiB  
Article
Pregnancy Vaccination with Gold Glyco-Nanoparticles Carrying Listeria monocytogenes Peptides Protects against Listeriosis and Brain- and Cutaneous-Associated Morbidities
by Ricardo Calderón-Gonzalez, Héctor Terán-Navarro, Elisabet Frande-Cabanes, Eva Ferrández-Fernández, Javier Freire, Soledad Penadés, Marco Marradi, Isabel García, Javier Gomez-Román, Sonsoles Yañez-Díaz and Carmen Álvarez-Domínguez
Nanomaterials 2016, 6(8), 151; https://doi.org/10.3390/nano6080151 - 19 Aug 2016
Cited by 29 | Viewed by 6592
Abstract
Listeriosis is a fatal infection for fetuses and newborns with two clinical main morbidities in the neonatal period, meningitis and diffused cutaneous lesions. In this study, we vaccinated pregnant females with two gold glyconanoparticles (GNP) loaded with two peptides, listeriolysin peptide 91–99 (LLO [...] Read more.
Listeriosis is a fatal infection for fetuses and newborns with two clinical main morbidities in the neonatal period, meningitis and diffused cutaneous lesions. In this study, we vaccinated pregnant females with two gold glyconanoparticles (GNP) loaded with two peptides, listeriolysin peptide 91–99 (LLO91–99) or glyceraldehyde-3-phosphate dehydrogenase 1–22 peptide (GAPDH1–22). Neonates born to vaccinated mothers were free of bacteria and healthy, while non-vaccinated mice presented clear brain affections and cutaneous diminishment of melanocytes. Therefore, these nanoparticle vaccines are effective measures to offer pregnant mothers at high risk of listeriosis interesting therapies that cross the placenta. Full article
(This article belongs to the Special Issue Nanoparticles in Immunology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>P0 corresponds to day 1 and P4 to day 4 of pups born to pregnant mothers. Groups of pregnant mothers were the following: non-vaccinated and challenged with <span class="html-italic">Listeria monocytogenes</span> (LM<sup>WT</sup>) (NV), vaccinated and challenged with LM<sup>WT</sup> (gold glyconanoparticles listeriolysin peptide 91–99 (GNP-LLO<sub>91–99</sub>) or gold glyconanoparticles glyceraldehyde-3-phosphate dehydrogenase 1–22 peptide (GNP-GAPDH<sub>1–22</sub>)) or non-vaccinated and non-infected with LM<sup>WT</sup> (control). P4-brain corresponds to brains of P4 neonates.</p>
Full article ">Figure 2
<p>(<b>a</b>) P4-brain image corresponds to mixed microglía (MG) cultures of cerebellum brains showing green fluorescence of green fluorescence protein <span class="html-italic">Listeria monocytogenes</span> (GFP-LM), red fluorescence of the macrophage marker F4/80 and blue fluoresecence of neuron tubuline. P4-MG corresponds to purified MG of mixed cultures showing green fluorescence of GFP-LM, red fluorescence of actin cytoskeleton stained with TRITC-phalloidin and blue fluorescence of tubuline; (<b>b</b>) colony-forming-units (CFU)/mL in MG of newborn P4 pups (NB) born to NV mothers, GNP-LLO<sub>91–99</sub> or GNP-GAPDH<sub>1–22</sub> vaccinated mothers. CFU/mL in livers and spleens of NV mothers are also examined; (<b>c</b>) Cytokines levels (pg/mL) in supernatants of MG of newborn P4 pups born to mothers as in <a href="#nanomaterials-06-00151-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
1619 KiB  
Review
Recent Prospects in the Inline Monitoring of Nanocomposites and Nanocoatings by Optical Technologies
by Elodie Bugnicourt, Timothy Kehoe, Marcos Latorre, Cristina Serrano, Séverine Philippe and Markus Schmid
Nanomaterials 2016, 6(8), 150; https://doi.org/10.3390/nano6080150 - 19 Aug 2016
Cited by 27 | Viewed by 7641
Abstract
Nanostructured materials have emerged as a key research field in order to confer materials with unique or enhanced properties. The performance of nanocomposites depends on a number of parameters, but the suitable dispersion of nanoparticles remains the key in order to obtain the [...] Read more.
Nanostructured materials have emerged as a key research field in order to confer materials with unique or enhanced properties. The performance of nanocomposites depends on a number of parameters, but the suitable dispersion of nanoparticles remains the key in order to obtain the full nanocomposites’ potential in terms of, e.g., flame retardance, mechanical, barrier, thermal properties, etc. Likewise, the performance of nanocoatings to obtain, for example, tailored surface affinity with selected liquids (e.g., for self-cleaning ability or anti-fog properties), protective effects against flame propagation, ultra violet (UV) radiation or gas permeation, is highly dependent on the nanocoating’s thickness and homogeneity. In terms of recent advances in the monitoring of nanocomposites and nanocoatings, this review discusses commonly-used offline characterization approaches, as well as promising inline systems. All in all, having good control over both the dispersion and thickness of these materials would help with reaching optimal and consistent properties to allow nanocomposites to extend their use. Full article
(This article belongs to the Special Issue Nanocomposite Coatings)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of the “tortuous pathway” created by the incorporation of exfoliated nanoplatelets into a polymer matrix film. Adapted from [<a href="#B23-nanomaterials-06-00150" class="html-bibr">23</a>].</p>
Full article ">Figure 2
<p>Scheme of different types of composite arising from the interaction of clays and polymers: (<b>a</b>) phase separated microcomposite; (<b>b</b>) intercalated nanocomposite; and (<b>c</b>) exfoliated nanocomposites. Adapted from [<a href="#B26-nanomaterials-06-00150" class="html-bibr">26</a>].</p>
Full article ">Figure 3
<p>Conventional characterization techniques to evaluate nanoparticles’ dispersion in different media.</p>
Full article ">
6229 KiB  
Article
Receptor-Meditated Endocytosis by Hyaluronic Acid@Superparamagnetic Nanovetor for Targeting of CD44-Overexpressing Tumor Cells
by Kwang Sik Yu, Meng Meng Lin, Hyun-Ju Lee, Ki-Sik Tae, Bo-Sun Kang, Je Hun Lee, Nam Seob Lee, Young Gil Jeong, Seung-Yun Han and Do Kyung Kim
Nanomaterials 2016, 6(8), 149; https://doi.org/10.3390/nano6080149 - 18 Aug 2016
Cited by 13 | Viewed by 5325
Abstract
The present report proposes a more rational hyaluronic acid (HA) conjugation protocol that can be used to modify the surface of the superparamagnetic iron oxide nanoparticles (SPIONs) by covalently binding the targeting molecules (HA) with glutamic acid as a molecular linker on peripheral [...] Read more.
The present report proposes a more rational hyaluronic acid (HA) conjugation protocol that can be used to modify the surface of the superparamagnetic iron oxide nanoparticles (SPIONs) by covalently binding the targeting molecules (HA) with glutamic acid as a molecular linker on peripheral surface of SPIONs. The synthesis of HA-Glutamic Acid (GA)@SPIONs was included oxidization of nanoparticle’s surface with H2O2 followed by activation of hydroxyl group and reacting glutamic acid as an intermediate molecule demonstrating transfection of lung cancer cells. Fourier transform infrared (FTIR) and zeta-potential studies confirmed the chemical bonding between amino acid linker and polysaccharides. 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) cytotoxicity assay showed that HA-SPIONs-treated cells remained 82.9% ± 2.7% alive at high particle dosage (200 µg/mL iron concentration), whereas GA-SPIONs and bare SPIONs (B-SPIONs) treated cells had only 59.3% ± 13.4% and 26.5% ± 3.1% survival rate at the same conditions, respectively. Confocal microscopy analysis showed increased cellular internalization of HA-SPIONs compared to non-interacting agarose coated SPIONs (AgA-SPIONs). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>–<b>c</b>) Three possible configurations of glutamic acid superparamagnetic iron oxide nanoparticles (GA-SPIONs); and (<b>d</b>) schematic illustration of hyaluronic acid superparamagnetic iron oxide nanoparticles (HA-SPIONs).</p>
Full article ">Figure 2
<p>Transmission electron microscopy (TEM) images of: (<b>a</b>) bare SPIONs (B-SPIONs); (<b>b</b>) hydroxyl group activated SPION (OH-SPIONs); (<b>c</b>) GA-SPIONs; and (<b>d</b>) HA-SPIONs.</p>
Full article ">Figure 3
<p>Fourier transform infrared spectra of B-SPIONs, OH-SPIONs, GA-SPIONs and HA-SPIONs in the wavenumber range of 2500 cm<sup>−1</sup> to 650 cm<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Zeta-potential of B-SPIONs, OH-SPIONs, GA-SPIONs, HA-SPIONs and agarose coated SPIONs (AgA-SPIONs).</p>
Full article ">Figure 5
<p>Histogram of cell viability after 24 h incubation with B-SPIONs (green), GA-SPIONs (red) and HA-SPIONs (blue) at 37 °C, 5% CO<sub>2</sub>. (* means the mean of the group is significantly different from the other two at the same Fe concentration at <span class="html-italic">p</span> = 0.05 level) (<span class="html-italic">n</span> ≥ 5).</p>
Full article ">Figure 6
<p>Confocal microscopy images of A459 cells only as control (<b>left</b> column); A459 cells incubated with 20 µg/mL fluorescein isothiocyanate (FITC)-HA-SPIONs (green colored) (<b>middle</b> column) and enlarged imaged of A459 cells with FITC-HA-SPIONs (<b>right</b> column) for: (<b>a</b>) 6 h; (<b>b</b>) 8 h; (<b>c</b>) 12 h; and (<b>d</b>) 24 h. Cells were stained with red membrane dye PKH26 prior to FTIC-HA-SPIONs incubation at 37 °C, 5% CO<sub>2</sub> (the scale bars of the left and middle column are 100 mm, and the scale bars of the right column are 20 mm).</p>
Full article ">Figure 6 Cont.
<p>Confocal microscopy images of A459 cells only as control (<b>left</b> column); A459 cells incubated with 20 µg/mL fluorescein isothiocyanate (FITC)-HA-SPIONs (green colored) (<b>middle</b> column) and enlarged imaged of A459 cells with FITC-HA-SPIONs (<b>right</b> column) for: (<b>a</b>) 6 h; (<b>b</b>) 8 h; (<b>c</b>) 12 h; and (<b>d</b>) 24 h. Cells were stained with red membrane dye PKH26 prior to FTIC-HA-SPIONs incubation at 37 °C, 5% CO<sub>2</sub> (the scale bars of the left and middle column are 100 mm, and the scale bars of the right column are 20 mm).</p>
Full article ">Figure 7
<p>Histogram of cellular iron uptake of A459 cells without any treatment (green) and those incubated with AgA-SPIONs (red) and HA-SPIONs (blue), cells with no particles treatment as control (green), at 1 h, 4 h, 8 h and 20 h incubation, at 37 °C, 5% CO<sub>2</sub>. (* means the mean of the group is significantly different from the other two groups at <span class="html-italic">p</span> = 0.05 level) (<span class="html-italic">n</span> ≥ 5).</p>
Full article ">
3844 KiB  
Article
Flexible Textile-Based Organic Transistors Using Graphene/Ag Nanoparticle Electrode
by Youn Kim, Yeon Ju Kwon, Kang Eun Lee, Youngseok Oh, Moon-Kwang Um, Dong Gi Seong and Jea Uk Lee
Nanomaterials 2016, 6(8), 147; https://doi.org/10.3390/nano6080147 - 16 Aug 2016
Cited by 19 | Viewed by 7392
Abstract
Highly flexible and electrically-conductive multifunctional textiles are desirable for use in wearable electronic applications. In this study, we fabricated multifunctional textile composites by vacuum filtration and wet-transfer of graphene oxide films on a flexible polyethylene terephthalate (PET) textile in association with embedding Ag [...] Read more.
Highly flexible and electrically-conductive multifunctional textiles are desirable for use in wearable electronic applications. In this study, we fabricated multifunctional textile composites by vacuum filtration and wet-transfer of graphene oxide films on a flexible polyethylene terephthalate (PET) textile in association with embedding Ag nanoparticles (AgNPs) to improve the electrical conductivity. A flexible organic transistor can be developed by direct transfer of a dielectric/semiconducting double layer on the graphene/AgNP textile composite, where the textile composite was used as both flexible substrate and conductive gate electrode. The thermal treatment of a textile-based transistor enhanced the electrical performance (mobility = 7.2 cm2·V−1·s−1, on/off current ratio = 4 × 105, and threshold voltage = −1.1 V) due to the improvement of interfacial properties between the conductive textile electrode and the ion-gel dielectric layer. Furthermore, the textile transistors exhibited highly stable device performance under extended bending conditions (with a bending radius down to 3 mm and repeated tests over 1000 cycles). We believe that our simple methods for the fabrication of graphene/AgNP textile composite for use in textile-type transistors can potentially be applied to the development of flexible large-area electronic clothes. Full article
(This article belongs to the Special Issue Textiles Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the vacuum filtration and subsequent transfer of graphene oxide (GO) film on a flexible textile substrate.</p>
Full article ">Figure 2
<p>Photographs of GO films on anodic aluminum oxide (AAO) membranes with different amounts of GO (<b>top</b>). Optical microscope images of GO films on polyethylene terephthalate (PET) textiles (<b>bottom</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Optical microscopy (OM) images and (<b>c</b>) scanning electron microscopy (SEM) image of graphene/AgNP textile composite.</p>
Full article ">Figure 4
<p>Schematic illustration of the fabrication process of the transistor device based on the graphene/silver nanoparticle (AgNP) textile composites. P3HT: Poly(3-hexylthiophene); S: Source electrode; D: Drain electrode.</p>
Full article ">Figure 5
<p>Photograph (<b>a</b>) and optical microscopy image (<b>b</b>) of textile-based transistor device.</p>
Full article ">Figure 6
<p>(<b>a</b>) <span class="html-italic">I</span><sub>D</sub>-<span class="html-italic">V</span><sub>D</sub> and (<b>b</b>) <span class="html-italic">I</span><sub>D</sub>-<span class="html-italic">V</span><sub>G</sub> characteristics of the textile-based transistor device.</p>
Full article ">Figure 7
<p><span class="html-italic">I</span><sub>D</sub>-<span class="html-italic">V</span><sub>G</sub> characteristics of the textile-based transistor device after thermal annealing at 80 °C for 15 min.</p>
Full article ">Figure 8
<p>Changes in the hole mobilities of the thermally-annealed transistor (blue square) and not-annealed transistor (red square) depending on the bending cycle.</p>
Full article ">
3686 KiB  
Article
Methanol-Tolerant Platinum-Palladium Catalyst Supported on Nitrogen-Doped Carbon Nanofiber for High Concentration Direct Methanol Fuel Cells
by Jiyoung Kim, Jin-Sung Jang, Dong-Hyun Peck, Byungrok Lee, Seong-Ho Yoon and Doo-Hwan Jung
Nanomaterials 2016, 6(8), 148; https://doi.org/10.3390/nano6080148 - 15 Aug 2016
Cited by 16 | Viewed by 5204
Abstract
Pt-Pd catalyst supported on nitrogen-doped carbon nanofiber (N-CNF) was prepared and evaluated as a cathode electrode of the direct methanol fuel cell (DMFC). The N-CNF, which was directly synthesized by the catalytic chemical vapor deposition from acetonitrile at 640 °C, was verified as [...] Read more.
Pt-Pd catalyst supported on nitrogen-doped carbon nanofiber (N-CNF) was prepared and evaluated as a cathode electrode of the direct methanol fuel cell (DMFC). The N-CNF, which was directly synthesized by the catalytic chemical vapor deposition from acetonitrile at 640 °C, was verified as having a change of electrochemical surface properties such as oxygen reduction reaction (ORR) activities and the electrochemical double layer compared with common carbon black (CB). To attain the competitive oxygen reduction reaction activity with methanol tolerance, the Pt and Pd metals were supported on the CB or the N-CNF. The physical and electrochemical characteristics of the N-CNF–supported Pt-Pd catalyst were examined and compared with catalyst supported on the CB. In addition, DMFC single cells using these catalysts as the cathode electrode were applied to obtain I-V polarization curves and constant current operating performances with high-concentration methanol as the fuel. Pt-Pd catalysts had obvious ORR activity even in the presence of methanol. The higher power density was obtained at all the methanol concentrations when it applied to the membrane electrode assembly (MEA) of the DMFC. When the N-CNF is used as the catalyst support material, a better performance with high-concentration methanol is expected. Full article
Show Figures

Figure 1

Figure 1
<p>The scanning electron microscope (SEM) and transmission electron microscope (TEM) images of the nitrogen doped carbon nanofiber (N-CNF).</p>
Full article ">Figure 2
<p>Presence of oxygen reduction reaction (ORR) activity in the N-CNF measured by conventional electrochemical equipment.</p>
Full article ">Figure 3
<p>Thermogravimetry (TG) profiles of commercial Pt catalyst and Pt-Pd binary catalysts on the carbon black (CB) or the N-CNF.</p>
Full article ">Figure 4
<p>Activity for ORR of the catalysts in 0.1 M HClO<sub>4</sub> with a rotating disk electrode at 1600 rpm.</p>
Full article ">Figure 5
<p>Activity for ORR of the catalysts in presence of 1 M methanol with oxygen.</p>
Full article ">Figure 6
<p>The direct methanol fuel cell (DMFC) unit-cell performances of the membrane electrode assemblies (MEAs) prepared with the catalysts of (<b>a</b>) commercial Pt/CB; (<b>b</b>) Pt<sub>1</sub>Pd<sub>4.1</sub>/CB and (<b>c</b>) Pt<sub>1</sub>Pd<sub>4.8</sub>/N-CNF with various methanol concentrations.</p>
Full article ">Figure 7
<p>Comparisons of power densities at (<b>a</b>) 0.4 V and (<b>b</b>) maximum peak point.</p>
Full article ">Figure 8
<p>Constant current operation of the MEAs with the catalysts as cathode materials.</p>
Full article ">
3893 KiB  
Review
Metallic Nanostructures Based on DNA Nanoshapes
by Boxuan Shen, Kosti Tapio, Veikko Linko, Mauri A. Kostiainen and Jari Jussi Toppari
Nanomaterials 2016, 6(8), 146; https://doi.org/10.3390/nano6080146 - 10 Aug 2016
Cited by 18 | Viewed by 8634
Abstract
Metallic nanostructures have inspired extensive research over several decades, particularly within the field of nanoelectronics and increasingly in plasmonics. Due to the limitations of conventional lithography methods, the development of bottom-up fabricated metallic nanostructures has become more and more in demand. The remarkable [...] Read more.
Metallic nanostructures have inspired extensive research over several decades, particularly within the field of nanoelectronics and increasingly in plasmonics. Due to the limitations of conventional lithography methods, the development of bottom-up fabricated metallic nanostructures has become more and more in demand. The remarkable development of DNA-based nanostructures has provided many successful methods and realizations for these needs, such as chemical DNA metallization via seeding or ionization, as well as DNA-guided lithography and casting of metallic nanoparticles by DNA molds. These methods offer high resolution, versatility and throughput and could enable the fabrication of arbitrarily-shaped structures with a 10-nm feature size, thus bringing novel applications into view. In this review, we cover the evolution of DNA-based metallic nanostructures, starting from the metallized double-stranded DNA for electronics and progress to sophisticated plasmonic structures based on DNA origami objects. Full article
(This article belongs to the Special Issue DNA-Based Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Atomic force microscope (AFM) image of silver nanowires [<a href="#B65-nanomaterials-06-00146" class="html-bibr">65</a>]; (<b>b</b>) continuous Pd nanowires using a 42-h incubation time and a 45 °C temperature [<a href="#B69-nanomaterials-06-00146" class="html-bibr">69</a>]; (<b>c</b>) when lowering the incubation time to 20 h, but using the same temperature, discontinuous Pd nanowires were formed [<a href="#B69-nanomaterials-06-00146" class="html-bibr">69</a>]; (<b>d</b>) continuous Pt nanowires formed by UV photoexcitation [<a href="#B83-nanomaterials-06-00146" class="html-bibr">83</a>]; (<b>e</b>) silver toroids formed by reducing silver salt on spermine treated ssDNA [<a href="#B77-nanomaterials-06-00146" class="html-bibr">77</a>]; (<b>f</b>) long Ni nanowires on lambda-DNA [<a href="#B72-nanomaterials-06-00146" class="html-bibr">72</a>]. (a) is reproduced with permission from [<a href="#B65-nanomaterials-06-00146" class="html-bibr">65</a>]. Copyright Nature Publishing Group, 1998; (b,c) are reproduced with permission from [<a href="#B69-nanomaterials-06-00146" class="html-bibr">69</a>]. Copyright John Wiley and Sons, 1998; (d) is reproduced with permission from [<a href="#B83-nanomaterials-06-00146" class="html-bibr">83</a>]. Copyright Elsevier, 2009; (e) is reproduced with permission from [<a href="#B77-nanomaterials-06-00146" class="html-bibr">77</a>]. Copyright American Chemical Society, 2005; (f) is reproduced with permission from [<a href="#B72-nanomaterials-06-00146" class="html-bibr">72</a>]. Copyright American Chemical Society, 2006.</p>
Full article ">Figure 2
<p>Upper panel (<b>a</b>–<b>c</b>) [<a href="#B23-nanomaterials-06-00146" class="html-bibr">23</a>]: (a) schematic of a 4 × 4 tile and nanoribbon assembly form from these tiles; (b) AFM image of a nanoribbon; (c) SEM image of a metallized silver nanoribbon, scale bar 500 nm; middle panel (<b>d</b>–<b>g</b>) [<a href="#B87-nanomaterials-06-00146" class="html-bibr">87</a>]: (<b>d</b>) scheme of a nanotube made of TX tiles; (e,f) TEM and SEM image of nanotubes; (g) SEM image of a metallized silver nanotube; scale bars in (e–g) are 100 nm, 1 μm and 1 μm, respectively; lower panel (<b>h</b>–<b>j</b>) [<a href="#B24-nanomaterials-06-00146" class="html-bibr">24</a>]: (h) assembly model of a nanotube from a single oligonucleotide with palindromic sequence, (i) AFM image of the nanotube; (<b>j</b>) metallized Pd nanotube. (a–c) are reproduced with permission from [<a href="#B23-nanomaterials-06-00146" class="html-bibr">23</a>]. Copyright The American Association for the Advancement of Science, 2003; (d–g) are reproduced with permission from [<a href="#B87-nanomaterials-06-00146" class="html-bibr">87</a>]. Copyright National Academy of Sciences, USA, 2004; (h–j) are reproduced with permission from [<a href="#B24-nanomaterials-06-00146" class="html-bibr">24</a>]. Copyright John Wiley and Sons, 2006.</p>
Full article ">Figure 3
<p>(<b>a</b>) AFM image of a Y-shaped DNA origami metallized with Au on mica. The origami shape before metallization is presented in the inset (scale bar 200 nm) [<a href="#B88-nanomaterials-06-00146" class="html-bibr">88</a>]. (<b>b</b>) DNA origami seeded with Pd<sup>2+</sup> and metallized with Au and the corresponding EDX results [<a href="#B89-nanomaterials-06-00146" class="html-bibr">89</a>]. (<b>c</b>,<b>d</b>) A circuit-like DNA origami metallized with Au (c) and Cu (d) with respective EDX as the inset [<a href="#B90-nanomaterials-06-00146" class="html-bibr">90</a>]. (a) is reproduced with permission from [<a href="#B88-nanomaterials-06-00146" class="html-bibr">88</a>]. Copyright American Chemical Society, 2011; (b) is reproduced with permission from [<a href="#B89-nanomaterials-06-00146" class="html-bibr">89</a>]. Copyright Royal Society of Chemistry, 2011; (c,d) are reproduced with permission from [<a href="#B90-nanomaterials-06-00146" class="html-bibr">90</a>]. Copyright American Chemical Society, 2013.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) Metallized H-shapes, two parallel bars and rings. Insets are the corresponding structures before metallization, scale bars are 250 nm [<a href="#B93-nanomaterials-06-00146" class="html-bibr">93</a>]. (<b>d</b>) SERS spectrum of aminobenzenethiol obtained by using the four-corner bound AuNPs structure (inset) for enhancement of the Raman signal. (<b>e</b>) Structures with just a single AuNP resulted in an insignificant SERS signal [<a href="#B94-nanomaterials-06-00146" class="html-bibr">94</a>]. (<b>f</b>) A single origami bar metallized with both gold and copper on either side [<a href="#B96-nanomaterials-06-00146" class="html-bibr">96</a>]. (<b>g</b>–<b>i</b>) T-shaped DNA structures with AuNPs bound to one branch, both branches (inset in (h)) and only on the edges of the T-branches (inset in (i)). The structures in the insets of (h) and (i) were further metallized into continuous metal structures as shown in (h) and (i). Scale bars in insets of (g) and (h) are 500 nm and of (i) 100 nm [<a href="#B91-nanomaterials-06-00146" class="html-bibr">91</a>]. (<b>j</b>,<b>k</b>) Ring and cross patterns fabricated by reducing gold seeds that were bound via electrostatic attraction [<a href="#B97-nanomaterials-06-00146" class="html-bibr">97</a>]. (a–c) are reproduced with permission from [<a href="#B93-nanomaterials-06-00146" class="html-bibr">93</a>]. Copyright American Chemical Society, 2011; (d–e) are reproduced with permission from [<a href="#B94-nanomaterials-06-00146" class="html-bibr">94</a>]. Copyright American Chemical Society, 2014; (f) is reproduced with permission from [<a href="#B96-nanomaterials-06-00146" class="html-bibr">96</a>]. Copyright American Chemical Society, 2014; (g–i) are reproduced with permission from [<a href="#B91-nanomaterials-06-00146" class="html-bibr">91</a>]. Copyright American Chemical Society, 2012; (j,k) are reproduced with permission from [<a href="#B97-nanomaterials-06-00146" class="html-bibr">97</a>]. Copyright John Wiley and Sons, 2011.</p>
Full article ">Figure 5
<p>(<b>a</b>) Top panel: casting metal particles with prescribed 3D shapes using programmable DNA nano-structure molds. Bottom panel: experimental results of the cast procedure. The scale bars are 20 nm [<a href="#B98-nanomaterials-06-00146" class="html-bibr">98</a>]. (<b>b</b>) Top panel: schematic views and TEM images of the AuNP seed grown inside the DNA origami mold without the lid. Bottom panel: side-by-side and head-to-tail designs. All scale bars correspond to 50 nm [<a href="#B99-nanomaterials-06-00146" class="html-bibr">99</a>]. (a) is reproduced with permission from [<a href="#B98-nanomaterials-06-00146" class="html-bibr">98</a>]. Copyright The American Association for the Advancement of Science, 2014; (b) is reproduced with permission from [<a href="#B99-nanomaterials-06-00146" class="html-bibr">99</a>]. Copyright American Chemical Society, 2014.</p>
Full article ">Figure 6
<p>(<b>a</b>) Method to transfer the negative pattern from the DNA nano-grid to a gold surface [<a href="#B100-nanomaterials-06-00146" class="html-bibr">100</a>]; (<b>b</b>) AFM image of the gold surface with the negative square grid from DNA assembly [<a href="#B100-nanomaterials-06-00146" class="html-bibr">100</a>]; (<b>c</b>) by using a dsDNA as a mask in an angled evaporation, an open area is formed on the evaporated film, which can be further utilized as an etching mask [<a href="#B101-nanomaterials-06-00146" class="html-bibr">101</a>]; (<b>d</b>) silver nanowire grown in the etched trench [<a href="#B101-nanomaterials-06-00146" class="html-bibr">101</a>]. (a,b) are reproduced with permission from [<a href="#B100-nanomaterials-06-00146" class="html-bibr">100</a>]. Copyright John Wiley and Sons, 2004; (c,d) are reproduced with permission from [<a href="#B101-nanomaterials-06-00146" class="html-bibr">101</a>]. Copyright John Wiley and Sons, 2007.</p>
Full article ">Figure 7
<p>(<b>a</b>) DNA origami-modulated etching of SiO<sub>2</sub> by HF vapor [<a href="#B102-nanomaterials-06-00146" class="html-bibr">102</a>]; (<b>b</b>) room-temperature CVD process for SiO<sub>2</sub> growth using DNA origami as a mask [<a href="#B104-nanomaterials-06-00146" class="html-bibr">104</a>]; (<b>c</b>) fabrication steps to produce high-resolution metallic shapes on the Si surface using DNA origami stencils [<a href="#B105-nanomaterials-06-00146" class="html-bibr">105</a>]; (<b>d</b>) similar cross-shaped structures fabricated from different metals via the same origami mask [<a href="#B105-nanomaterials-06-00146" class="html-bibr">105</a>]. (a) is reproduced with permission from [<a href="#B102-nanomaterials-06-00146" class="html-bibr">102</a>]. Copyright American Chemical Society, 2011; (b) is reproduced with permission from [<a href="#B104-nanomaterials-06-00146" class="html-bibr">104</a>]. Copyright American Chemical Society, 2013; (c,d) are reproduced with permission from [<a href="#B105-nanomaterials-06-00146" class="html-bibr">105</a>]. Copyright Royal Society of Chemistry, 2015.</p>
Full article ">
148 KiB  
Editorial
Nanostructured Solar Cells
by Guanying Chen, Zhijun Ning and Hans Ågren
Nanomaterials 2016, 6(8), 145; https://doi.org/10.3390/nano6080145 - 9 Aug 2016
Cited by 11 | Viewed by 5971
Abstract
We are glad to announce the Special Issue “Nanostructured Solar Cells”, published in Nanomaterials. This issue consists of eight articles, two communications, and one review paper, covering major important aspects of nanostructured solar cells of varying types. From fundamental physicochemical investigations to technological [...] Read more.
We are glad to announce the Special Issue “Nanostructured Solar Cells”, published in Nanomaterials. This issue consists of eight articles, two communications, and one review paper, covering major important aspects of nanostructured solar cells of varying types. From fundamental physicochemical investigations to technological advances, and from single junction solar cells (silicon solar cell, dye sensitized solar cell, quantum dots sensitized solar cell, and small molecule organic solar cell) to tandem multi-junction solar cells, all aspects are included and discussed in this issue to advance the use of nanotechnology to improve the performance of solar cells with reduced fabrication costs. Full article
(This article belongs to the Special Issue Nanostructured Solar Cells)
1585 KiB  
Communication
Silicon Nanowire Photocathodes for Photoelectrochemical Hydrogen Production
by Soundarrajan Chandrasekaran, Thomas Nann and Nicolas H. Voelcker
Nanomaterials 2016, 6(8), 144; https://doi.org/10.3390/nano6080144 - 5 Aug 2016
Cited by 14 | Viewed by 5543
Abstract
The performance of silicon for water oxidation and hydrogen production can be improved by exploiting the antireflective properties of nanostructured silicon substrates. In this work, silicon nanowires were fabricated by metal-assisted electroless etching of silicon. An enhanced photocurrent density of −17 mA/cm2 [...] Read more.
The performance of silicon for water oxidation and hydrogen production can be improved by exploiting the antireflective properties of nanostructured silicon substrates. In this work, silicon nanowires were fabricated by metal-assisted electroless etching of silicon. An enhanced photocurrent density of −17 mA/cm2 was observed for the silicon nanowires coated with an iron sulphur carbonyl catalyst when compared to bare silicon nanowires (−5 mA/cm2). A substantial amount of 315 µmol/h hydrogen gas was produced at low bias potentials for the silicon nanowires coated with an iron sulphur carbonyl catalyst. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Scanning electron microscopy (SEM) image of the fabricated silicon nanowires (SiNWs) for 10 min of etching time. The inset shows the cross-sectional SEM image; (<b>B</b>) transmission electron microscopy (TEM) image of an individual SiNW.</p>
Full article ">Figure 2
<p>Current density measurements for the planar silicon (<b>A</b>), bare SiNWs (<b>B</b>), planar silicon coated with Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub> catalyst (<b>C</b>) and SiNWs coated with Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub> catalyst (<b>D</b>).</p>
Full article ">Figure 3
<p>A long run measurement of planar silicon (black trace) and SiNW (red trace) coated with Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub> catalyst respectively in 0.1 M H<sub>2</sub>SO<sub>4</sub> at a bias potential of −500 mV over 7 h (under illumination).</p>
Full article ">Figure 4
<p>Gas chromatography (GC) analysis of the sample gas (500 µL) from the headspace of the photoelectrochemical (PEC) cell for planar silicon (black trace) and SiNW (red trace) coated with Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub> catalyst, respectively.</p>
Full article ">
2748 KiB  
Article
Distribution of Iron Oxide Core-Titanium Dioxide Shell Nanoparticles in VX2 Tumor Bearing Rabbits Introduced by Two Different Delivery Modalities
by Tamer Refaat, Derek West, Samar El Achy, Vamsi Parimi, Jasmine May, Lun Xin, Kathleen R. Harris, William Liu, Michael Beau Wanzer, Lydia Finney, Evan Maxey, Stefan Vogt, Reed A. Omary, Daniele Procissi, Andrew C. Larson, Tatjana Paunesku and Gayle E. Woloschak
Nanomaterials 2016, 6(8), 143; https://doi.org/10.3390/nano6080143 - 3 Aug 2016
Cited by 7 | Viewed by 6339
Abstract
This work compares intravenous (IV) versus fluoroscopy-guided transarterial intra-catheter (IC) delivery of iron oxide core-titanium dioxide shell nanoparticles (NPs) in vivo in VX2 model of liver cancer in rabbits. NPs coated with glucose and decorated with a peptide sequence from cortactin were administered [...] Read more.
This work compares intravenous (IV) versus fluoroscopy-guided transarterial intra-catheter (IC) delivery of iron oxide core-titanium dioxide shell nanoparticles (NPs) in vivo in VX2 model of liver cancer in rabbits. NPs coated with glucose and decorated with a peptide sequence from cortactin were administered to animals with developed VX2 liver cancer. Two hours after NPs delivery tumors, normal liver, kidney, lung and spleen tissues were harvested and used for a series on histological and elemental analysis tests. Quantification of NPs in tissues was done both by bulk inductively coupled plasma mass spectrometry (ICP-MS) analysis and by hard X-ray fluorescence microscopy. Both IV and IC NPs injection are feasible modalities for delivering NPs to VX2 liver tumors with comparable tumor accumulation. It is possible that this is an outcome of the fact that VX2 tumors are highly vascularized and hemorrhagic, and therefore enhanced permeability and retention (EPR) plays the most significant role in accumulation of nanoparticles in tumor tissue. It is, however, interesting to note that IV delivery led to increased sequestration of NPs by spleen and normal liver tissue, while IC delivery lead to more NP positive Kupffer cells. This difference is most likely a direct outcome of blood flow dynamics. Armed with this knowledge about nanoparticle delivery, we plan to test them as radiosensitizers in the future. Full article
(This article belongs to the Special Issue Nanomaterials for Cancer Therapies)
Show Figures

Figure 1

Figure 1
<p>Intravenous (IV) and transarterial intra-catheter (IC) injected rabbits’ VX2 tumors and surrounding liver tissue stained by H&amp;E, histochemical (HC) stain (detecting the nanoparticles) and Ki67 Antibodies.</p>
Full article ">Figure 2
<p>Details from <a href="#app1-nanomaterials-06-00143" class="html-app">Supplemental Figure S3</a>. (<b>a</b>) IV injected rabbit’s lung showing metastatic nodular tumor deposits disposed in a peripheral location composed of poorly differentiated squamoid cells expanding the alveolar septae (100× H&amp;E); (<b>b</b>) A photomicrograph of the IV injected rabbit lung showing alveolar septae expanded and infiltrated by inflammatory infiltrate, predominantly eosinophils (200× H&amp;E); (<b>c</b>) IC rabbit lung showing marked thickening and destruction of alveolar spetae by metastatic deposits and inflammatory infiltrate composed mainly of granulocytes (arrowheads in inset, 400×). Blood vessels are engorged with granulocytes (arrow head) (100×); (<b>d</b>,<b>e</b>) IV injected rabbit’s splenic red pulp displaying dilated and congested sinusoids. Arrows point to follicles (100× H&amp;E); (<b>f</b>) IC injected rabbit spleen showing congested and expanded red pulp with influx of macrophages and proliferated Littoral cells; (<b>g</b>) IV injected rabbit renal cortex showing mild to moderate cloudy changes in the tubules (100× H&amp;E); (<b>h</b>) IC injected rabbit kidney showing cortical tubules exhibiting severe cloudy swelling and vesicular changes (100× H&amp;E); (<b>i</b>) An IV injected rabbit liver micrograph displaying mild lobular disarray with hepatocyte ballooning (200× H&amp;E); (<b>j</b>) IC injected rabbit’s liver showing dilated congested central veins and sinusoids. Hepatocytes show ballooning and microvesicular steatosis (200× H&amp;E); (<b>k</b>) Photomicrograph showing part of VX2 tumor mass with a central necrotic cavity (100× H&amp;E); (<b>l</b>) A photomicrograph of VX2 tumor showing infiltrating malignant cells on the left side, tumor infiltrating lymphocytes on the right, and a dilated sinusoid with tumor thrombus in the center (arrow) (100× H&amp;E); (<b>m</b>) Photomicrograph of VX2 tumor showing nests of squamous cell carcinoma (yellow arrow), surrounded by a peripheral zone of tumor infiltrating lymphocytes (red arrow), compressed hepatocytes and fibroblasts; (<b>n</b>) Control rabbit lung micrograph showing multiple microscopic metastatic tumor deposits (arrows), congested microvasculature and patent alveoli (100× H&amp;E); (<b>o</b>) Control Rabbit spleen micrograph showing white pulp follicles with narrow intervening red pulp sinusoids, minimal congestion ascribed to manipulation during necropsy; (<b>p</b>) Control rabbit kidney micrograph showing patent cortical tubules and unremarkable glomerular tufts with absence of congested vessels (100× H&amp;E); (<b>q</b>) Photomicrograph of control rabbit liver (100× H&amp;E); (<b>r</b>) Control Rabbit micrograph of VX2 liver tumor showing malignant squamoid cells (arrow) surrounding a central necrotic cavity (star) (100× H&amp;E).</p>
Full article ">Figure 3
<p>Details from <a href="#app1-nanomaterials-06-00143" class="html-app">Supplemental Figure S3</a>. Histochemical Staining for NPs: (<b>a</b>) IV injected rabbit’s VX2 tumors showing weak uneven staining of tumor masses. Arrow shows weak staining of surrounding hepatocytes (100×); (<b>b</b>) Photomicrograph showing nuclear staining of IV injected rabbit’s VX2 tumor cells with weak cytoplasmic stain. Surrounding tissues are not stained (200×); (<b>c</b>) Intense brown nuclear and cytoplasmic staining of the IC injected rabbit’s tumor masses (100×); (<b>d</b>) Micrograph of IC injected rabbit’s VX2 tumor showing intense positive nuclear and cytoplasmic staining of cells (400×); (<b>e</b>) Photomicrograph of control rabbit VX2 showing faint non-specific staining of malignant tumor cells (arrow) and necrotic tissues (star); (<b>f</b>) Micrograph of IV injected rabbit’s spleen showing positively stained red pulp macrophages and marginal zone macrophages (arrows) (200×); (<b>g</b>) Micrograph of IC injected rabbit’s spleen showing positively stained red pulp macrophages (400×); (<b>h</b>) Micrograph of control rabbit’s spleen red pulp showing absence of positive staining (200×); (<b>i</b>) Positively stained Kupffer cell in a sinusoid of an IV injected rabbit (400×); (<b>j</b>) IC injected rabbit’s micrographs showing intense staining of periportal hepatocytes and bile duct epithelium (arrow) (200×); (<b>k</b>) Micrograph of control liver showing no positive staining (200×); (<b>l</b>) IV injected rabbit renal medullary tubules showing intense positive cytoplasmic stain (200×); (<b>m</b>) IC injected rabbit’s micrograph of rabbit medullary tubules of kidney showing intense positive cytoplasmic stain (200×); (<b>n</b>) Control rabbit’s kidney micrograph showing no positive staining (100×); (<b>o</b>) IV injected rabbit’s lung showing a nodular metastatic deposit with positively stained tumor cells (200×); (<b>p</b>) IV injected rabbit’s lung displaying positively stained tumor metastatic deposits, in addition to alveolar macrophages. Arrow is pointing at a positively stained alveolar macrophage (200×); (<b>q</b>) IC rabbit lung showing positive brown staining of metastatic deposits within the alveolar septae (red arrowhead) and the bronchiolar epithelial lining cells (black arrowhead) (200×); (<b>r</b>) Micrograph of control rabbit’s lung showing absence of positive staining in metastatic deposits (arrows) and alveolar macrophages (100×).</p>
Full article ">Figure 4
<p>(<b>a</b>) NP positive Kupffer cells and (<b>b</b>) total number of Kupffer cells in livers of intravenous (IV) and intra-catheter (IC) injected and control rabbits. <span class="html-italic">Y</span>-axis indicates total number of Kupffer cells per field of view, averaged from 20 such areas, images acquired at 400× (for example, see one such area of view in <a href="#nanomaterials-06-00143-f003" class="html-fig">Figure 3</a>i).</p>
Full article ">Figure 5
<p>Overview of XFM images for NP distribution after intravenous (IV) and transarterial intra-catheter (IC) injections into control rabbits VX2 tumor and liver, kidney, lung, and spleen tissues. The mapping for the Titanium (<b>left</b>) and Zinc (<b>right</b>) concentrations and content informs about tissue shape, and potentially even its health status. It should be noted that these are false color images showing distribution of elemental concentrations from the lowest (<b>black to brown</b>) to highest (<b>yellow to white</b>) within the elemental concentrations distribution of each sample (see color scale bar in lower right depicting spectra of colors matching lowest concentration (in a given sample)—black) to highest (<b>white</b>) across a “red temperature” scale. Thus, samples with very low elemental concentrations for a given element (e.g., Ti in control samples) show a “salt and pepper” pixel distribution indicating that background signal levels predominate in the sample. It should be noted that paraffin embedding disrupts distribution of free ions such as potassium, while sulfur and zinc distributions still well represent a tissue outline because Zn finger proteins are accumulated in each cells’ nucleus [<a href="#B14-nanomaterials-06-00143" class="html-bibr">14</a>,<a href="#B15-nanomaterials-06-00143" class="html-bibr">15</a>,<a href="#B16-nanomaterials-06-00143" class="html-bibr">16</a>,<a href="#B17-nanomaterials-06-00143" class="html-bibr">17</a>,<a href="#B18-nanomaterials-06-00143" class="html-bibr">18</a>,<a href="#B19-nanomaterials-06-00143" class="html-bibr">19</a>]. Distribution of nanoparticles, much larger and insoluble <span class="html-italic">in situ</span>, is considered to be accurately depicted by XFM [<a href="#B14-nanomaterials-06-00143" class="html-bibr">14</a>,<a href="#B15-nanomaterials-06-00143" class="html-bibr">15</a>,<a href="#B16-nanomaterials-06-00143" class="html-bibr">16</a>,<a href="#B17-nanomaterials-06-00143" class="html-bibr">17</a>,<a href="#B18-nanomaterials-06-00143" class="html-bibr">18</a>,<a href="#B19-nanomaterials-06-00143" class="html-bibr">19</a>].</p>
Full article ">Figure 6
<p>Graphic representation of data in <a href="#nanomaterials-06-00143-t001" class="html-table">Table 1</a>. (<b>Top</b>) Ti concentration (in femtograms per pixel) in different tissues from nanoparticle treated animals and controls; data for each animal are shown. (<b>Bottom</b>) Mean ratio of Ti (femtograms) vs. S (femtograms) per pixel; please note that data structure allows comparison on a per pixel basis. Thus, information provided by the bottom graph allows us to monitor the situation most accurately, e.g., if tissue integrity is uneven, resultant uneven concentration of nanoparticles in tissue will not be misinterpreted. Likewise, false background (“salt and pepper” signal pattern for Ti) is effectively removed when false signals are divided by real signal for S.</p>
Full article ">
3083 KiB  
Review
Review of Recent Developments on Using an Off-Lattice Monte Carlo Approach to Predict the Effective Thermal Conductivity of Composite Systems with Complex Structures
by Feng Gong, Hai M. Duong and Dimitrios V. Papavassiliou
Nanomaterials 2016, 6(8), 142; https://doi.org/10.3390/nano6080142 - 30 Jul 2016
Cited by 17 | Viewed by 5571
Abstract
Here, we present a review of recent developments for an off-lattice Monte Carlo approach used to investigate the thermal transport properties of multiphase composites with complex structure. The thermal energy was quantified by a large number of randomly moving thermal walkers. Different modes [...] Read more.
Here, we present a review of recent developments for an off-lattice Monte Carlo approach used to investigate the thermal transport properties of multiphase composites with complex structure. The thermal energy was quantified by a large number of randomly moving thermal walkers. Different modes of heat conduction were modeled in appropriate ways. The diffusive heat conduction in the polymer matrix was modeled with random Brownian motion of thermal walkers within the polymer, and the ballistic heat transfer within the carbon nanotubes (CNTs) was modeled by assigning infinite speed of thermal walkers in the CNTs. Three case studies were conducted to validate the developed approach, including three-phase single-walled CNTs/tungsten disulfide (WS2)/(poly(ether ether ketone) (PEEK) composites, single-walled CNT/WS2/PEEK composites with the CNTs clustered in bundles, and complex graphene/poly(methyl methacrylate) (PMMA) composites. In all cases, resistance to heat transfer due to nanoscale phenomena was also modeled. By quantitatively studying the influencing factors on the thermal transport properties of the multiphase composites, it was found that the orientation, aggregation and morphology of fillers, as well as the interfacial thermal resistance at filler-matrix interfaces would limit the transfer of heat in the composites. These quantitative findings may be applied in the design and synthesis of multiphase composites with specific thermal transport properties. Full article
(This article belongs to the Special Issue Computational Modeling and Simulations of Carbon Nanomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic plot of the SWNT/WS<sub>2</sub>/PEEK model: (<b>a</b>) a WS<sub>2</sub> nanoparticle (110 nm diameter) is placed in the center of a PEEK cube with a side length of 925 nm, while 317 SWNTs (2 nm diameter and 500 nm length) are randomly distributed in the PEEK cube. The WS<sub>2</sub> particle is painted red and the nanotubes are black in the figure. Constant heat flux is applied along the <span class="html-italic">x</span> direction by creating a hot surface and a cooled surface (Reproduced with permission from [<a href="#B18-nanomaterials-06-00142" class="html-bibr">18</a>]. Copyright Elsevier, 2015); (<b>b</b>) a contour plot of thermal walker distribution in the center <span class="html-italic">xy</span> plane of the SWNT/WS<sub>2</sub>/PEEK model at the thermal steady state (Reproduced with permission from [<a href="#B24-nanomaterials-06-00142" class="html-bibr">24</a>]. Copyright American Chemical Society, 2015).</p>
Full article ">Figure 2
<p>Validation of the developed approach by comparing the simulation results with the experimental data from Reference [<a href="#B9-nanomaterials-06-00142" class="html-bibr">9</a>]. The side lengths of the PEEK cubes in 0.1/0.9/99.0, 0.5/0.5/99.0 and 0.9/0.1/99.0 compositions were 760, 925 and 1580 nm, respectively. The interfacial thermal resistance at SWNT-PEEK interface was used as 1.0 × 10<sup>−8</sup> m<sup>2</sup>·K/W for Nan et al.’s effective medium theory (EMT) [<a href="#B32-nanomaterials-06-00142" class="html-bibr">32</a>]. The error bars represent the standard deviation of the results obtained from 3 separate simulations with different distribution of SWNTs. Reproduced with permission from [<a href="#B18-nanomaterials-06-00142" class="html-bibr">18</a>]. Copyright Elsevier, 2014.</p>
Full article ">Figure 3
<p>Effects of the interfacial thermal resistances at (<b>a</b>) SWNT-PEEK and (<b>b</b>) WS<sub>2</sub>-PEEK interfaces on the <span class="html-italic">K</span><sub>eff</sub> of SWNT/WS<sub>2</sub>/PEEK composites. The 0.5/0.5/99.0 composition was used for this quantitative study. The models with different SWNT orientation (e.g., SWNTs parallel to the heat flux, SWNTs randomly orientated to the heat flux, and SWNTs perpendicular to the heat flux) were built to study the effect of SWNT orientations. The error bars represent the standard deviation of the results from 3 separate simulations with different distribution of SWNTs. Reproduced with permission from [<a href="#B18-nanomaterials-06-00142" class="html-bibr">18</a>]. Copyright Elsevier, 2014.</p>
Full article ">Figure 4
<p>Effects of (<b>a</b>) length and (<b>b</b>) diameter of SWNTs on the <span class="html-italic">K</span><sub>eff</sub> of SWNT/WS<sub>2</sub>/PEEK composites. The length was varied from 100 to 900 nm, corresponding to an aspect ratio from 50 to 450. The diameter was varied from 2 to 8 nm while the length was kept as 500 nm. Different compositions with randomly orientated SWNTs were chosen to study the effect of SWNT diameter. The error bars represent the standard deviation of the results obtained from 3 separate simulations with different distribution of SWNTs. Reproduced with permission from [<a href="#B18-nanomaterials-06-00142" class="html-bibr">18</a>]. Copyright Elsevier, 2014.</p>
Full article ">Figure 5
<p>Schematic plot of the SWNT/WS<sub>2</sub>/PEEK model with SWNT bundles: (<b>a</b>) A WS<sub>2</sub> nanoparticle (painted red) with 110 nm diameter is located in the center of a PEEK cube (925 × 925 × 925 nm<sup>3</sup>). A total of 960 SWNTs (2 nm diameter and 500 nm length, painted black) are randomly dispersed in the model, forming 45 bundles with 20 SWNTs in each bundle and 60 unbundled SWNTs. Constant heat flux is applied along × direction; (<b>b</b>) composite with individual SWNTs and SWNT bundles oriented parallel to the heat-flux direction. Reproduced with permission from [<a href="#B24-nanomaterials-06-00142" class="html-bibr">24</a>]. Copyright American Chemical Society, 2015.</p>
Full article ">Figure 6
<p>Effects of the morphology of SWNT bundles on the <span class="html-italic">K</span><sub>eff</sub> of the SWNT/WS<sub>2</sub>/PEEK composites: (<b>a</b>) bundle number; and (<b>b</b>) the number of individual SWNTs in per bundle. The results for SWNTs with different orientations (parallel, random and perpendicular to the heat flux direction) are all presented. The error bars represent the standard deviation of the results obtained from 3 separate simulations with different distribution of SWNTs and SWNT bundles. Reproduced with permission from [<a href="#B24-nanomaterials-06-00142" class="html-bibr">24</a>]. Copyright American Chemical Society, 2015.</p>
Full article ">Figure 7
<p>Effect of the SWNT-SWNT thermal resistance on the <span class="html-italic">K</span><sub>eff</sub> of the SWNT/WS<sub>2</sub>/PEEK composites with 12–48 SWNT bundles. The SWNTs were randomly distributed in the composites. The individual SWNT number in each bundle was kept at 20. The critical TBR (dashed line) was estimated to be <span class="html-italic">R</span><sub>c</sub> = 0.155 × 10<sup>−8</sup> m<sup>2</sup>·K/W by intersecting the <span class="html-italic">K</span><sub>eff</sub> curves of different SWNT bundles, as shown in the insert figure. Reproduced with permission from [<a href="#B24-nanomaterials-06-00142" class="html-bibr">24</a>]. Copyright American Chemical Society, 2015.</p>
Full article ">Figure 8
<p>(<b>a</b>) Validation of the developed graphene/PMMA model by comparing the simulation results with the experimental data for different volume fractions of graphene sheets (0.67%, 1.34%, 2.01% and 2.50%). The insert figure is the set-up scheme of the comparative infrared microscopy technique for measuring the thermal conductivity. More details of the experimental set-up can be found in Reference [<a href="#B4-nanomaterials-06-00142" class="html-bibr">4</a>]; (<b>b</b>) the thermal conductivity of GA-PMMA composites as a function of graphene volume fraction. The interfacial thermal resistance between graphene sheets and PMMA was estimated to be <span class="html-italic">R</span><sub>bd</sub> = 1.906 × 10<sup>−8</sup> m<sup>2</sup>·K/W in the developed model. The same value was utilized in the composites with parallel- and perpendicular-oriented graphene. In the modified EMT, the utilized thermal conductivity of graphene and the <span class="html-italic">R</span><sub>bd</sub> of graphene-PMMA were 100 W/m·K and 1.0 × 10<sup>−8</sup> m<sup>2</sup>·K/W, respectively. The error bars represent the standard deviation of the results from separate measurements of thermal conductivity of graphene/PMMA composites. Reproduced with permission from [<a href="#B4-nanomaterials-06-00142" class="html-bibr">4</a>]. Copyright Elsevier, 2015.</p>
Full article ">
3301 KiB  
Article
Galactosylated Liposomes for Targeted Co-Delivery of Doxorubicin/Vimentin siRNA to Hepatocellular Carcinoma
by Hea Ry Oh, Hyun-Young Jo, James S. Park, Dong-Eun Kim, Je-Yoel Cho, Pyung-Hwan Kim and Keun-Sik Kim
Nanomaterials 2016, 6(8), 141; https://doi.org/10.3390/nano6080141 - 30 Jul 2016
Cited by 77 | Viewed by 8125
Abstract
The combination of therapeutic nucleic acids and chemotherapeutic drugs has shown great promise for cancer therapy. In this study, asialoglycoprotein receptors (ASGPR) targeting-ligand-based liposomes were tested to determine whether they can co-deliver vimentin siRNA and doxorubicin to hepatocellular carcinoma (HCC) selectively. To achieve [...] Read more.
The combination of therapeutic nucleic acids and chemotherapeutic drugs has shown great promise for cancer therapy. In this study, asialoglycoprotein receptors (ASGPR) targeting-ligand-based liposomes were tested to determine whether they can co-deliver vimentin siRNA and doxorubicin to hepatocellular carcinoma (HCC) selectively. To achieve this goal, we developed an ASGPR receptor targeted co-delivery system called gal-doxorubicin/vimentin siRNA liposome (Gal-DOX/siRNA-L). The Gal-DOX/siRNA-L was created via electrostatic interaction of galactose linked-cationic liposomal doxorubicin (Gal-DOX-L) on vimentin siRNA. Previous studies have shown that Gal-DOX/siRNA-L inhibited tumor growth by combined effect of DOX and vimentin siRNA than single delivery of either DOX or vimentin siRNA. These Gal-DOX/siRNA-Ls showed stronger affinity to human hepatocellular carcinoma cells (Huh7) than other cells (lung epithelial carcinoma, A549). These liposomes also have demonstrated that novel hepatic drug/gene delivery systems composed of cationic lipid (DMKE: O,O’-dimyristyl-N-lysyl glutamate), cholesterol, galactosylated ceramide, POPC (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine), and PEG2000-DSPE (distearoyl phosphatidyl ethanolamine) at 2:1:1:1:0.2 (moral ratios) can be used as an effective drug/gene carrier specifically targeting the liver in vivo. These results suggest that Gal-DOX-siRNA-L could effectively target tumor cells, enhance transfection efficacy and subsequently achieve the co-delivery of DOX and siRNA, demonstrating great potential for synergistic anti-tumor therapy. Full article
(This article belongs to the Special Issue Nanomaterials for Tissue Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Gel retardation analysis and fluorescence intensities measurement of Gal-DOX/siRNA-L complexes at various N/P ratios. (<b>a</b>) Gel retardation assay of different formulations of siRNA, containing FITC-siRNA (200 nM) per sample, on a 2% TAE (Tris Acetate EDTA) agarose gel at 100 V for 30 min, then investigated by UV (Ultraviolet) illuminator; and (<b>b</b>) fluorescence intensities measured uncomplexed free FITC-labeled siRNA at varying N/P ratios in the gel by a FUSION SL chemiluminescence analyzer and software (VILBER, Suarlée, Belgium). (* <span class="html-italic">p</span> &lt; 0.05 vs. N/P = 6; mean ± SD; <span class="html-italic">n</span> = 3 replicates/group).</p>
Full article ">Figure 2
<p>Specific binding of the Gal-DOX-Ls to human hepatocarcinoma Huh7 cells. (<b>a</b>) A549 and Huh7 cells were incubated with either plane liposomes (DOTAP/Chol liposomes) or Gal-DOX-Ls for 15 min in culture media. These liposomes having the rhodamine-conjugated DOPE lipids (red color) were observed on the cell surface by a fluorescence microscope (JuLi-Smart Fluorescence Cell Imager, NanoEnTek Inc., Seoul, South Korea). (<b>b</b>) For asialofetuin (AF) inhibition, AF in culture media added as increasing concentration for 30 min before the treated of Gal liposome (20 μg). Cells were analyzed after 30 min last treatment of Gal liposome. Red fluorescence color in each cells were investigated by a Tali image-based cytometer (Thermo Fisher Scientific Inc., Waltham, MA, USA). (<b>c</b>) These liposomes having the rhodamine-conjugated DOPE lipids were observed on the cell surface by an Axio Zeiss A1 Imager compound microscope (Carl Zeiss, Oberkochen, Germany). Error bars represent standard deviation of three independent experiments. * <span class="html-italic">p</span> &lt; 0.01 when compared to with no treatment (asialofetuin, 0 mg).</p>
Full article ">Figure 3
<p>Cytotoxic effect of Gal-DOX-L formation to Huh7 cells. Cell viability was measured after A549 cells and Huh7 cells were incubated with various formations of DOX tor 48 h. The cells were treated with free DOX and DOX-encapsulated in Gal-liposomes (Gal-DOX-L), where each formation was adjusted to contain various amount of DOX. As a control liposome, cells were also treated with liposomal doxorubicin (DOX-L [0.2 μM]) lacking galactose. No treatment means that cells were not treated. Data are represented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 compared with Free DOX (IC<sub>50</sub>) treatment.</p>
Full article ">Figure 4
<p>Cell uptake and cytotoxicity by Gal-lipoplexes-mediated siRNA. (<b>a</b>) Internalization of FITC-siRNA by Gal-lipoplexes in the ASGPR-expressing Huh7 cells via receptor-ligand mediated endocytosis. A549 cells and Huh7 cells were incubated with Gal-lipoplexes of a various FITC-siRNA concentration (50, 100, and 200 pmole) for 2 h in culture media. FITC-labeled siRNAs were visualized in green at the ASGPR-expressing Huh7 cells; (<b>b</b>) inhibition of the expression of vimentin in Huh7 cells by Gal-lipoplexes (vimentin siRNA). Huh7 cells were treated with control free siRNA or Gal-lipoplexes (vimentin siRNA, 50–200 pmole). Total cell lysates (40 μg) of Huh7 cells were subjected to SDS-PAGE (sodium dodecyl sulfate polyacrylamide gel electrophoresis). The separated proteins were analyzed by Western blot method to detect vimentin as described. The β-actin detection was included as a loading control. Values of relative protein expression are expressed as the mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. Gal-lipoplexes [vimentin siRNA, 100 pmole]. (<b>c</b>) To measure in vitro cytotoxicity of Gal-lipoplexes (vimentin siRNA), Huh 7 cells and A549 cells were incubated with free vimentin siRNA (200 pmole), lipoplexes lacking of galactose (vimentin siRNA, 200 pmole), or Gal-lipoplexes (vimentin siRNA) containing various concentration of vimentin siRNA (50–200 pmole) for 48 h. Cell viability was each monitored by the MTT assay method using EZ-CyTox reagents after 48 h. Data are represented as the mean ± standard deviation (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs. Gal-lipoplexes (vimentin siRNA, 100 pmole) in Huh7 cells; ** <span class="html-italic">p</span> &lt; 0.01 vs. Gal-lipoplexes (vimentin siRNA, 200 pmole) in A549 cells.</p>
Full article ">Figure 5
<p>Cytotoxic effects of Gal-DOX/siRNA-L in Huh7 cells. The control group was treated with saline. To measure the cytotoxicity synergic effect of Gal-DOX/siRNA-L, Huh 7 cells were incubated with various formation containing with the final vimentin siRNA concentration of 200 pmole and DOX concentration of 0.2 μM. Cell viability was each monitored by the MTT assay method using EZ-CyTox reagents after 48 h. Data are represented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.01 vs. Gal-DOX-L (0.2 μM); ** <span class="html-italic">p</span> &lt; 0.05 vs. Gal-lipoplexes (siRNA, 200 pmole).</p>
Full article ">Figure 6
<p>Biodistribution and in vivo anti-tumor efficacy of Gal-DOX/siRNA-L in xenograft nude mice. (<b>a</b>) The mice were injected via tail vein at a single dose of free DOX (5 mg/kg), DOX/siRNA-L (5 mg/kg DOX, 150 μg/kg siRNA), and Gal-DOX/siRNA-L (5 mg/kg DOX, 150 μg/kg siRNA) (<span class="html-italic">n</span> = 4) when the tumors grew to about 400 mm<sup>3</sup>. The tissues were collected at 4 h after the injections. Date are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 4); and (<b>b</b>) growth curves of xenograft tumors treated with saline, free DOX, DOX/siRNA-L, Gal-DOX-L, Gal-Lipoplexes (siRNA), and Gal-DOX/siRNA-L by intravenous injection once a week for four weeks (5 mg/kg DOX or/and 150 μg/kg vimentin siRNA). The curves present the changes of tumor sizes from the day of injection (day 0). Results are expressed as mean and standard deviation (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05, significant compared with Gal-DOX-L or Gal-lipoplexes (siRNA).</p>
Full article ">Scheme 1
<p>Schematic illustration of formation of Gal-DOX/siRNA-L (galactosylated-doxorubicin/ vimentin siRNA liposome). Firstly, doxorubicin (DOX) was encapsulated in cationic galactosylated liposomes by the pH-gradient insertion method. Subsequently, galactosylated liposomal DOX (Gal-DOX-L) and siRNA were co-loaded (Gal-DOX/siRNA complexes) by electrostatic attraction.</p>
Full article ">
5288 KiB  
Article
Optoelectronic and Electrochemical Properties of Vanadium Pentoxide Nanowires Synthesized by Vapor-Solid Process
by Ko-Ying Pan and Da-Hua Wei
Nanomaterials 2016, 6(8), 140; https://doi.org/10.3390/nano6080140 - 29 Jul 2016
Cited by 34 | Viewed by 6872
Abstract
Substantial synthetic vanadium pentoxide (V2O5) nanowires were successfully produced by a vapor-solid (VS) method of thermal evaporation without using precursors as nucleation sites for single crystalline V2O5 nanowires with a (110) growth plane. The micromorphology and [...] Read more.
Substantial synthetic vanadium pentoxide (V2O5) nanowires were successfully produced by a vapor-solid (VS) method of thermal evaporation without using precursors as nucleation sites for single crystalline V2O5 nanowires with a (110) growth plane. The micromorphology and microstructure of V2O5 nanowires were analyzed by scanning electron microscope (SEM), energy-dispersive X-ray spectroscope (EDS), transmission electron microscope (TEM) and X-ray diffraction (XRD). The spiral growth mechanism of V2O5 nanowires in the VS process is proved by a TEM image. The photo-luminescence (PL) spectrum of V2O5 nanowires shows intrinsic (410 nm and 560 nm) and defect-related (710 nm) emissions, which are ascribable to the bound of inter-band transitions (V 3d conduction band to O 2p valence band). The electrical resistivity could be evaluated as 64.62 Ω·cm via four-point probe method. The potential differences between oxidation peak and reduction peak are 0.861 V and 0.470 V for the first and 10th cycle, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram illustrating the thermal evaporation set-up with three main processes including evaporation, reaction and condensation.</p>
Full article ">Figure 2
<p>(<b>a</b>) Scanning electron microscope (SEM) image of V<sub>2</sub>O<sub>5</sub> nanowires; (<b>b</b>) Enlarged SEM image of V<sub>2</sub>O<sub>5</sub> nanowires.</p>
Full article ">Figure 3
<p>Energy-dispersive X-ray spectrometer (EDS) analysis at spot 1 of V<sub>2</sub>O<sub>5</sub> nanowires. Inset: SEM image of spot 1.</p>
Full article ">Figure 4
<p>X-ray diffraction (XRD) pattern of V<sub>2</sub>O<sub>5</sub> nanowires.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>)Transmission electron microscope (TEM) images of a V<sub>2</sub>O<sub>5</sub> nanowire. Upper left inset of (<b>a</b>): high-resolution transmission electron microscope (HRTEM) image. Lower right inset of (<b>a</b>): selected-area-electron-diffraction (SAED) pattern.</p>
Full article ">Figure 6
<p>Schematic diagram illustrating the periodic bond chain (PBC) theory. Flat surfaces (F-face) with one PBC penetrate {100} planes. Stepped surfaces (S-face), {110} planes, own two PBCs. Kinked surfaces (K-face), {111} planes, have three PBCs.</p>
Full article ">Figure 7
<p>(<b>a</b>) SEM image taken from the sample prepared only at the beginning of the growth of V<sub>2</sub>O<sub>5</sub> nanowires by vapor-solid (VS) process; (<b>b</b>) The enlarged SEM image.</p>
Full article ">Figure 8
<p>Photo-luminescence (PL) spectrum of V<sub>2</sub>O<sub>5</sub> nanowires.</p>
Full article ">Figure 9
<p>I-V curve (the red curve) of a single V<sub>2</sub>O<sub>5</sub> nanowire measured by four-probe method. Lower right inset is the SEM image of the device fabricated by focused ion beam system (FIB).</p>
Full article ">Figure 10
<p>Cyclic voltammetry (CV) curve of V<sub>2</sub>O<sub>5</sub> nanowires in 2 M KCl at 50 m·Vs<sup>−1</sup> scan rate. The red curve recorded at the first cycle and the blue dash curve recorded during the 10th cycle, respectively.</p>
Full article ">
2132 KiB  
Review
DNA-Based Enzyme Reactors and Systems
by Veikko Linko, Sami Nummelin, Laura Aarnos, Kosti Tapio, J. Jussi Toppari and Mauri A. Kostiainen
Nanomaterials 2016, 6(8), 139; https://doi.org/10.3390/nano6080139 - 27 Jul 2016
Cited by 63 | Viewed by 9233
Abstract
During recent years, the possibility to create custom biocompatible nanoshapes using DNA as a building material has rapidly emerged. Further, these rationally designed DNA structures could be exploited in positioning pivotal molecules, such as enzymes, with nanometer-level precision. This feature could be used [...] Read more.
During recent years, the possibility to create custom biocompatible nanoshapes using DNA as a building material has rapidly emerged. Further, these rationally designed DNA structures could be exploited in positioning pivotal molecules, such as enzymes, with nanometer-level precision. This feature could be used in the fabrication of artificial biochemical machinery that is able to mimic the complex reactions found in living cells. Currently, DNA-enzyme hybrids can be used to control (multi-enzyme) cascade reactions and to regulate the enzyme functions and the reaction pathways. Moreover, sophisticated DNA structures can be utilized in encapsulating active enzymes and delivering the molecular cargo into cells. In this review, we focus on the latest enzyme systems based on novel DNA nanostructures: enzyme reactors, regulatory devices and carriers that can find uses in various biotechnological and nanomedical applications. Full article
(This article belongs to the Special Issue DNA-Based Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic view of an enzymatic nanoreactor built from DNA ([S] = substrate, [P] = product). By taking advantage of the high addressability and modularity of the DNA nanostructures, enzymes can be attached and arranged with nanometer-scale precision. As an example, glucose oxidase (GOx, purple)–horseradish peroxidase (HRP, green) cascade pairs have been assembled into a confined reaction space provided by two tubular DNA origami nanostructures (orange and yellow cages).</p>
Full article ">Figure 2
<p>(<b>a</b>) A DNA origami technique. A long scaffold strand is folded into a desired shape with the help of short staple strands [<a href="#B14-nanomaterials-06-00139" class="html-bibr">14</a>]; (<b>b</b>) Multilayer DNA origami in square and honeycomb lattice [<a href="#B15-nanomaterials-06-00139" class="html-bibr">15</a>,<a href="#B16-nanomaterials-06-00139" class="html-bibr">16</a>]; (<b>c</b>) DNA origami with curvatures and bends [<a href="#B17-nanomaterials-06-00139" class="html-bibr">17</a>,<a href="#B18-nanomaterials-06-00139" class="html-bibr">18</a>]; (<b>d</b>) Scaffold-free fabrication of DNA nanoshapes. Numerous target shapes can be fabricated by selecting subsets of strands from the cubic-like ‘molecular canvas’ [<a href="#B19-nanomaterials-06-00139" class="html-bibr">19</a>]; (<b>e</b>) A fully automated top-down design method to create meshed DNA origami structures [<a href="#B21-nanomaterials-06-00139" class="html-bibr">21</a>]; (<b>f</b>) DNA origami structures can be glued together by taking advantage of the blunt-end stacking and the shape-complementarity of the origami units [<a href="#B22-nanomaterials-06-00139" class="html-bibr">22</a>]. (<b>a</b>) is reproduced with permission from [<a href="#B14-nanomaterials-06-00139" class="html-bibr">14</a>]. Copyright Nature Publishing Group, 2006. (<b>b</b>) is reproduced with permission from [<a href="#B16-nanomaterials-06-00139" class="html-bibr">16</a>]. Copyright Nature Publishing Group, 2011. A sphere in (<b>c</b>) is reproduced with permission from [<a href="#B17-nanomaterials-06-00139" class="html-bibr">17</a>]. Copyright The American Association for the Advancement of Science, 2011. A gear-like object in (<b>c</b>) is reproduced with permission from [<a href="#B18-nanomaterials-06-00139" class="html-bibr">18</a>]. Copyright The American Association for the Advancement of Science, 2009. (<b>d</b>) is reproduced with permission from [<a href="#B19-nanomaterials-06-00139" class="html-bibr">19</a>]. Copyright The American Association for the Advancement of Science, 2012. (<b>e</b>) is reproduced with permission from [<a href="#B21-nanomaterials-06-00139" class="html-bibr">21</a>]. Copyright The American Association for the Advancement of Science, 2016. (<b>f</b>) is reproduced with permission from [<a href="#B22-nanomaterials-06-00139" class="html-bibr">22</a>]. Copyright The American Association for the Advancement of Science, 2015.</p>
Full article ">Figure 3
<p>(<b>a</b>) A glucose oxidase (GOx) – horseradish peroxidase (HRP) enzyme cascade pair assembled on a rectangular DNA origami [<a href="#B43-nanomaterials-06-00139" class="html-bibr">43</a>]; (<b>b</b>) A rectangular DNA origami shape with attached enzyme cascade pairs (GOx and HRP) can be rolled into tubular shapes [<a href="#B44-nanomaterials-06-00139" class="html-bibr">44</a>]; (<b>c</b>) A swinging arm for cofactor transfer between the enzymes (malate dehydrogenase (MDH) and glucose-6-phosphate dehydrogenase (G6pDH)) assembled on a DNA tile [<a href="#B45-nanomaterials-06-00139" class="html-bibr">45</a>]; (<b>d</b>) A modular and tubular DNA origami-based enzyme cascade (GOx and HRP) nanoreactor [<a href="#B46-nanomaterials-06-00139" class="html-bibr">46</a>]; (<b>e</b>) An artifical enzyme cascade (xylose reductase (XR) and xylitol dehydrogenase (XDR)) performing a cofactor coupled cascade reaction on DNA origami [<a href="#B47-nanomaterials-06-00139" class="html-bibr">47</a>]; (<b>f</b>) An artificial three-enzyme (lactate dehydrogenase (LDH), MDH and oxaloacetate decarboxylase (OAD)) pathway organized using a DNA nanostructure [<a href="#B48-nanomaterials-06-00139" class="html-bibr">48</a>]. (<b>a</b>) is reproduced with permission from [<a href="#B43-nanomaterials-06-00139" class="html-bibr">43</a>]. Copyright American Chemical Society, 2012. (<b>b</b>) is reproduced with permission from [<a href="#B44-nanomaterials-06-00139" class="html-bibr">44</a>]. Copyright American Chemical Society, 2013. (<b>c</b>) is reproduced with permission from [<a href="#B45-nanomaterials-06-00139" class="html-bibr">45</a>]. Copyright Nature Publishing Group, 2014. (<b>d</b>) is reproduced with permission from [<a href="#B46-nanomaterials-06-00139" class="html-bibr">46</a>]. Published by The Royal Society of Chemistry, 2015. (<b>e</b>) is reproduced with permission from [<a href="#B47-nanomaterials-06-00139" class="html-bibr">47</a>]. Copyright American Chemical Society, 2016. (<b>f</b>) is reproduced with permission from [<a href="#B48-nanomaterials-06-00139" class="html-bibr">48</a>]. Copyright John Wiley and Sons, 2016.</p>
Full article ">Figure 4
<p>(<b>a</b>) Nanotweezers to regulate enzyme activity [<a href="#B49-nanomaterials-06-00139" class="html-bibr">49</a>]; (<b>b</b>) Tubular nanoreactor with switchable lid to control the flowthrough of the reaction compounds [<a href="#B53-nanomaterials-06-00139" class="html-bibr">53</a>]; (<b>c</b>) DNA origami nanoactuator that can be driven by, e.g., single-stranded DNA (ssDNA) strands or restriction enzymes [<a href="#B54-nanomaterials-06-00139" class="html-bibr">54</a>]; (<b>d</b>) Aptamer-based logical molecular circuit to control thrombin activity [<a href="#B55-nanomaterials-06-00139" class="html-bibr">55</a>]. (<b>a</b>) is reproduced with permission from [<a href="#B49-nanomaterials-06-00139" class="html-bibr">49</a>] Copyright Nature Publishing Group, 2013. (<b>b</b>) is reproduced with permission from [<a href="#B53-nanomaterials-06-00139" class="html-bibr">53</a>]. Copyright The Royal Society of Chemistry, 2016. (<b>c</b>) is reproduced with permission from [<a href="#B54-nanomaterials-06-00139" class="html-bibr">54</a>]. Published by Nature Publishing Group, 2016. (<b>d</b>) is reproduced with permission from [<a href="#B55-nanomaterials-06-00139" class="html-bibr">55</a>]. Copyright American Chemical Society, 2012.</p>
Full article ">Figure 5
<p>(<b>a</b>) Hollow DNA origami containers: a box with a switchable lid [<a href="#B56-nanomaterials-06-00139" class="html-bibr">56</a>] and a tetrahedron [<a href="#B57-nanomaterials-06-00139" class="html-bibr">57</a>]; (<b>b</b>) Two half-cages equipped with glucose oxidase (GOx, orange) and horseradish peroxidase (HRP, green) can form a closed box and encapsulate the enzymes. The closed box efficiently protects the assembled enzyme cascade pair from protease digestion [<a href="#B59-nanomaterials-06-00139" class="html-bibr">59</a>]; (<b>c</b>) A DNA origami-based nanocarrier loaded with luciferase (LUC) enzymes. The enzyme activity can be tuned by coating the carrier with cationic polymers [<a href="#B60-nanomaterials-06-00139" class="html-bibr">60</a>]; (<b>d</b>) A light-triggered release of proteins such as bovine serum albumin (BSA) from DNA origami container [<a href="#B61-nanomaterials-06-00139" class="html-bibr">61</a>]; (<b>e</b>) A DNA cage that can trap and release an enzyme (HRP) through the temperature-controlled conformational changes [<a href="#B62-nanomaterials-06-00139" class="html-bibr">62</a>]; (<b>f</b>) β-galactosidase (β-gal) can be coated by DNA strands for significantly enhanced cellular delivery [<a href="#B63-nanomaterials-06-00139" class="html-bibr">63</a>]. A box with a lid in (<b>a</b>) is reproduced with permission from [<a href="#B56-nanomaterials-06-00139" class="html-bibr">56</a>]. Copyright Nature Publishing Group, 2009. A tetrahedron in (<b>a</b>) is reproduced with permission from [<a href="#B57-nanomaterials-06-00139" class="html-bibr">57</a>]. Copyright American Chemical Society, 2009. (<b>b</b>) is reproduced with permission from [<a href="#B59-nanomaterials-06-00139" class="html-bibr">59</a>]. Published by Nature Publishing Group, 2016. (<b>c</b>) is reproduced with permission from [<a href="#B60-nanomaterials-06-00139" class="html-bibr">60</a>]. Copyright The Royal Society of Chemistry, 2016. (<b>d</b>) is reproduced with permission from [<a href="#B61-nanomaterials-06-00139" class="html-bibr">61</a>]. Copyright American Chemical Society, 2016. (<b>e</b>) is reproduced with permission from [<a href="#B62-nanomaterials-06-00139" class="html-bibr">62</a>]. Copyright American Chemical Society, 2013. (<b>f</b>) is reproduced with permission from [<a href="#B63-nanomaterials-06-00139" class="html-bibr">63</a>]. Copyright American Chemical Society, 2015.</p>
Full article ">
3721 KiB  
Article
Synthesis of p-Co3O4/n-TiO2 Nanoparticles for Overall Water Splitting under Visible Light Irradiation
by Qiang Zhang, Zhenyin Hai, Aoqun Jian, Hongyan Xu, Chenyang Xue and Shengbo Sang
Nanomaterials 2016, 6(8), 138; https://doi.org/10.3390/nano6080138 - 27 Jul 2016
Cited by 29 | Viewed by 7348
Abstract
p-Co3O4/n-TiO2 nanoparticles (~400 nm) for photocatalysis were prepared via carbon assisted method and sol-gel method in this work. The paper also studied the application of visible light illuminated p-Co3O4/n-TiO2 nanocomposites cocatalyst to the [...] Read more.
p-Co3O4/n-TiO2 nanoparticles (~400 nm) for photocatalysis were prepared via carbon assisted method and sol-gel method in this work. The paper also studied the application of visible light illuminated p-Co3O4/n-TiO2 nanocomposites cocatalyst to the overall pure water splitting into H2 and O2. In addition, the H2 evolution rate of the p-Co3O4/n-TiO2 nanocomposites is 25% higher than that of the pure Co3O4 nanoparticles. Besides, according to the results of the characterizations, the scheme of visible light photocatalytic water splitting is proposed, the Co3O4 of the nanocomposites is excited by visible light, and the photo-generated electrons and holes existing on the conduction band of Co3O4 and valence band of TiO2 have endowed the photocatalytic evolution of H2 and O2 with higher efficiency. The optimal evolution rate of H2 and O2 is 8.16 μmol/h·g and 4.0 μmol/h·g, respectively. Full article
(This article belongs to the Special Issue Nanoscale in Photocatalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diagram of visible light water splitting system.</p>
Full article ">Figure 2
<p>X-ray diffraction (XRD) pattern of the: (<b>a</b>) Co<sub>3</sub>O<sub>4</sub> sample; (<b>b</b>) TiO<sub>2</sub> sample; and (<b>c</b>) TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> sample. a.u.: any unit.</p>
Full article ">Figure 3
<p>X-ray photoelectron spectroscopy (XPS) spectra of the TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> sample.</p>
Full article ">Figure 4
<p>Transmission electron microscopy (TEM) images of: (<b>a</b>) the Co<sub>3</sub>O<sub>4</sub> sample; (<b>b</b>) TiO<sub>2</sub> sample; and (<b>c</b>) TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> sample. High Resolution Transmission Electron Microscopy (HRTEM) images of: (<b>d</b>) the Co<sub>3</sub>O<sub>4</sub> sample; (<b>e</b>) TiO<sub>2</sub> sample; and (<b>f</b>) TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> sample.</p>
Full article ">Figure 5
<p>(<b>a</b>) Nitrogen adsorption–desorption isotherms; and (<b>b</b>) Barrett-Joyner-Halenda (BJH) pore size distributions of TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> sample.</p>
Full article ">Figure 6
<p>(<b>a</b>) Ultraviolet–visible (UV-Vis) spectra; and (<b>b</b>) Valence-band XPS spectra of TiO<sub>2</sub> and Co<sub>3</sub>O<sub>4</sub> nanoparticles. (Ahv)<sup>2</sup>–hv curve of: (<b>c</b>) Co<sub>3</sub>O<sub>4</sub> nanoparticles; and (<b>d</b>) TiO<sub>2</sub> nanoparticles. (<b>e</b>) Schematic diagram of the water splitting reaction of the TiO<sub>2</sub>-Co<sub>3</sub>O<sub>4</sub> heterostructures.</p>
Full article ">Figure 7
<p>Photocatalytic H<sub>2</sub> evolution on Co<sub>3</sub>O<sub>4</sub> nanocomposites under visible-light irradiation using 0.02 g photocatalyst suspended in 200 mL pure water solution in a Pyrex glass cell.</p>
Full article ">Figure 8
<p>(<b>a</b>) Photocatalytic H<sub>2</sub> and O<sub>2</sub> evolution on Co<sub>3</sub>O<sub>4</sub>-TiO<sub>2</sub> nanocomposites under visible-light irradiation using 0.02 g photocatalyst suspended in 200 mL pure water solution in a Pyrex glass cell; and (<b>b</b>) cycling measurements of H<sub>2</sub> and O<sub>2</sub> evolution through direct photocatalytic water splitting with Co<sub>3</sub>O<sub>4</sub>-TiO<sub>2</sub> nanocomposites under visible light.</p>
Full article ">
5121 KiB  
Article
Green Synthesis of Hierarchically Structured Silver-Polymer Nanocomposites with Antibacterial Activity
by María Jesús Hortigüela, Luis Yuste, Fernando Rojo and Inmaculada Aranaz
Nanomaterials 2016, 6(8), 137; https://doi.org/10.3390/nano6080137 - 25 Jul 2016
Cited by 13 | Viewed by 6365
Abstract
The in situ formation of silver nanoparticles (AgNPs) aided by chondroitin sulfate and the preparation of a hierarchically structured silver-polymer nanocomposite with antimicrobial activity is shown. Green synthesis of AgNPs is carried out by thermal treatment (80 and 90 °C) or UV irradiation [...] Read more.
The in situ formation of silver nanoparticles (AgNPs) aided by chondroitin sulfate and the preparation of a hierarchically structured silver-polymer nanocomposite with antimicrobial activity is shown. Green synthesis of AgNPs is carried out by thermal treatment (80 and 90 °C) or UV irradiation of a chondroitin sulfate solution containing AgNO3 without using any further reducing agents or stabilizers. Best control of the AgNPs size and polydispersity was achieved by UV irradiation. The ice-segregation-induced self-assembly (ISISA) process, in which the polymer solution containing the AgNPs is frozen unidirectionally, and successively freeze-drying were employed to produce the chondroitin sulfate 3D scaffolds. The scaffolds were further crosslinked with hexamethylene diisocyanate vapors to avoid water solubility of the 3D structures in aqueous environments. The antimicrobial activity of the scaffolds was tested against Escherichia coli. The minimum inhibitory concentration (MIC) found for AgNPs-CS (chondroitin sulfate) scaffolds was ca. 6 ppm. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of chondroitin-4-sulfate, where R<sub>1</sub> = H; R<sub>2</sub> = SO<sub>3</sub>H; R<sub>3</sub> = H.</p>
Full article ">Figure 2
<p>UV-Vis spectra of silver-containing chondroitin sulphate solutions (CS 10 wt. % and 0.16 mM AgNO<sub>3</sub>) after thermal treatment at 80 °C (<b>A</b>) and 90 °C (<b>B</b>) and UV-treatment (<b>C</b>). Thermal treatments were performed over 0, 90, 120, 210 and 270 min (lines a, b, c, d and e, respectively). UV treatments were performed over 120, 180 and 240 min (lines a, b and c). Sample irradiated during 240 min rested 24 h before measuring the exciton peak (line d).</p>
Full article ">Figure 3
<p>Transmission electron microscopy (TEM) micrographs of AgNPs-CS produced by thermal treatment at 80 °C (<b>A</b>); UV irradiation of fresh sample (<b>B</b>); and sample rested for 24 h (<b>C</b>). Bars are 50 nm in panel (<b>A</b>) and 20 nm in panels (<b>B</b>) and (<b>C</b>). CS: chondroitin sulfate.</p>
Full article ">Figure 4
<p>Particle size distribution estimated from TEM micrographs of AgNPs produced by UV treatment (<b>A</b>) and by thermal treatment (<b>B</b>). <span class="html-italic">N</span> =100.</p>
Full article ">Figure 5
<p>Metallic nature of AgNPs. Diffraction pattern (<b>A</b>) and X-ray photoelectron spectroscopy (XPS) spectrum (<b>B</b>) of the AgNPs produced by UV irradiation.</p>
Full article ">Figure 6
<p>Scanning electron microscopy (SEM) micrograph of AgNPs-CS crosslinked scaffold. Bar 50 µm.</p>
Full article ">Figure 7
<p>Silver release from AgNPs-CS crosslinked scaffolds in phosphate saline buffer (PSB) at 37 °C. <span class="html-italic">N</span> =2.</p>
Full article ">Figure 8
<p><span class="html-italic">E. coli</span> growth inhibition for different Ag concentrations. Ag 0 ppm (■); Ag 1.32 ppm (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#595959"> <mo mathsize="small">■</mo> </mstyle> </mrow> </semantics> </math>); Ag 3.76 ppm (<math display="inline"> <semantics> <mrow> <mstyle mathcolor="#AEAAAA"> <mo mathsize="small">■</mo> </mstyle> </mrow> </semantics> </math>); Ag 6.26 ppm (□).</p>
Full article ">Figure 9
<p>Antimicrobial activity of CS crosslinked scaffold (<b>A</b>); AgNPs-CS crosslinked scaffold (<b>B</b>); AgNO<sub>3</sub> crosslinked scaffold (<b>C</b>); AgNO<sub>3</sub> aqueous solution (Ag<sup>+</sup>: 6.25 ppm) (<b>D</b>), and AgNO<sub>3</sub> aqueous solution (Ag<sup>+</sup>: 12.5 ppm).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop