[go: up one dir, main page]

Next Issue
Volume 5, June
Previous Issue
Volume 4, December
 
 

Nanomaterials, Volume 5, Issue 1 (March 2015) – 23 articles , Pages 1-385

  • 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:
4903 KiB  
Article
Unidirectional Wave Propagation in Low-Symmetric Colloidal Photonic-Crystal Heterostructures
by Vassilios Yannopapas
Nanomaterials 2015, 5(1), 376-385; https://doi.org/10.3390/nano5010376 - 19 Mar 2015
Cited by 9 | Viewed by 5572
Abstract
We show theoretically that photonic crystals consisting of colloidal spheres exhibit unidirectional wave propagation and one-way frequency band gaps without breaking time-reversal symmetry via, e.g., the application of an external magnetic field or the use of nonlinear materials. Namely, photonic crystals with low [...] Read more.
We show theoretically that photonic crystals consisting of colloidal spheres exhibit unidirectional wave propagation and one-way frequency band gaps without breaking time-reversal symmetry via, e.g., the application of an external magnetic field or the use of nonlinear materials. Namely, photonic crystals with low symmetry such as the monoclinic crystal type considered here as well as with unit cells formed by the heterostructure of different photonic crystals show significant unidirectional electromagnetic response. In particular, we show that the use of scatterers with low refractive-index contrast favors the formation of unidirectional frequency gaps which is the optimal route for achieving unidirectional wave propagation. Full article
(This article belongs to the Special Issue Nanophotonic Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Unit cell of 3D PC with one-way photonic band gaps: monoclinic crystal consisting of four non-primitive planes of spheres parallel to the (001) surface at positions (0, 0, 0), (0, 0, <span class="html-italic">a</span>/2), (0, 0, <span class="html-italic">a</span>), and (−0.3<span class="html-italic">a</span>, 0, 3<span class="html-italic">a</span>/2).</p>
Full article ">Figure 2
<p>Finite slab of the photonic crystal of <a href="#nanomaterials-05-00376-f002" class="html-fig">Figure 2</a> consisting of eight unit layers.</p>
Full article ">Figure 3
<p>Transmittance for s- (a) and p- (c) polarized light incident with <b>k</b><sub>||</sub> = (0.25, 0)π/<span class="html-italic">a</span> on a finite slab of the PC (see <a href="#nanomaterials-05-00376-f002" class="html-fig">Figure 2</a>) consisting of eight unit layers whereas type-A spheres we have considered silica SiO<sub>2</sub> (ϵ = 2.1) spheres, as type-B polystyrene (ϵ = 2.6) spheres and as type-C silicon (ϵ = 11.9) (Si) spheres. <span class="html-italic">T</span><sup>+</sup> (<span class="html-italic">T</span><sup>−</sup>) is the transmittance for light incident from the left (right) (001) faces of the slab. (b) Frequency band structure of the infinitely periodic PC of <a href="#nanomaterials-05-00376-f001" class="html-fig">Figure 1</a> for <b>k</b><sub>||</sub>= (0.25, 0)<span class="html-italic">π/a</span>.</p>
Full article ">Figure 4
<p>Transmittance for s- (<b>a</b>) and p- (<b>c</b>) polarized light incident with <b>k</b><sub>||</sub> = (0.25, 0)π/<span class="html-italic">a</span> on a finite slab of the photonic crystal (PC) (see <a href="#nanomaterials-05-00376-f002" class="html-fig">Figure 2</a>) consisting of eight unit layers whereas type-A spheres we have considered silica SiO<sub>2</sub> (ϵ = 2.1) spheres, as type-B germanium (ϵ = 16.2) spheres and as type-C silicon (ϵ = 11.9) (Si) spheres. <span class="html-italic">T</span><sup>+</sup> (<span class="html-italic">T</span><sup>−</sup>) is the transmittance for light incident from the left (right) (001) faces of the slab; (<b>b</b>) Frequency band structure of the infinitely periodic PC of <a href="#nanomaterials-05-00376-f001" class="html-fig">Figure 1</a> for <b>k</b><sub>||</sub> = (0.25, 0)π/<span class="html-italic">a</span>.</p>
Full article ">Figure 5
<p>Transmittance for s- (<b>a</b>) and p- (<b>c</b>) polarized light incident with <b>k</b><sub>||</sub> = (0.25, 0)π/<span class="html-italic">a</span> on a finite slab of the PC of <a href="#nanomaterials-05-00376-f001" class="html-fig">Figure 1</a> consisting of eight unit layers whereas type-A spheres we have considered silica SiO<sub>2</sub> (ϵ = 2.1) spheres, as type-B polystyrene (ϵ = 2.6) spheres and as type-C sapphire (ϵ = 3.13) (Si) spheres. <span class="html-italic">T</span><sup>+</sup> (<span class="html-italic">T</span><sup>−</sup>) is the transmittance for light incident from the left (right) (001) faces of the slab; (<b>b</b>) Frequency band structure of the infinitely periodic PC of <a href="#nanomaterials-05-00376-f001" class="html-fig">Figure 1</a> for <b>k</b><sub>||</sub> = (0.25, 0)π/<span class="html-italic">a</span>.</p>
Full article ">
2640 KiB  
Article
On the Mass Fractal Character of Si-Based Structural Networks in Amorphous Polymer Derived Ceramics
by Sabyasachi Sen and Scarlett Widgeon
Nanomaterials 2015, 5(1), 366-375; https://doi.org/10.3390/nano5010366 - 17 Mar 2015
Cited by 12 | Viewed by 5403
Abstract
The intermediate-range packing of SiNxC4−x (0 ≤ x ≤ 4) tetrahedra in polysilycarbodiimide and polysilazane-derived amorphous SiCN ceramics is investigated using 29Si spin-lattice relaxation nuclear magnetic resonance (SLR NMR) spectroscopy. The SiCN network in the polysilylcarbodiimide-derived ceramic consists [...] Read more.
The intermediate-range packing of SiNxC4−x (0 ≤ x ≤ 4) tetrahedra in polysilycarbodiimide and polysilazane-derived amorphous SiCN ceramics is investigated using 29Si spin-lattice relaxation nuclear magnetic resonance (SLR NMR) spectroscopy. The SiCN network in the polysilylcarbodiimide-derived ceramic consists predominantly of SiN4 tetrahedra that are characterized by a 3-dimensional spatial distribution signifying compact packing of such units to form amorphous Si3N4 clusters. On the other hand, the SiCN network of the polysilazane-derived ceramic is characterized by mixed bonded SiNxC4−x tetrahedra that are inefficiently packed with a mass fractal dimension of Df ~2.5 that is significantly lower than the embedding Euclidean dimension (D = 3). This result unequivocally confirms the hypothesis that the presence of dissimilar atoms, namely, 4-coordinated C and 3-coordinated N, in the nearest neighbor environment of Si along with some exclusion in connectivity between SiCxN4−x tetrahedra with widely different N:C ratios and the absence of bonding between C and N result in steric hindrance to an efficient packing of these structural units. It is noted that similar inefficiencies in packing are observed in polymer-derived amorphous SiOC ceramics as well as in proteins and binary hard sphere systems. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>29</sup>Si MAS nuclear magnetic resonance (NMR) spectra of GM35 (<b>bottom</b>) and PMVS (<b>top</b>) samples. Spinning sidebands are denoted by asterisks. Structural assignments of the peaks to different Si coordination environments are indicated with arrows.</p>
Full article ">Figure 2
<p>Double logarithmic plots of the recovery of <sup>29</sup>Si magnetization after saturation, plotted as a function of the delay time for GM35 and PMVS samples. Lines are the linear least-squares fits to the data and represent a power law recovery of magnetization (see text for details) over a time span covering more than three orders of magnitude. The slopes of these lines are reported on the plot.</p>
Full article ">Figure 3
<p>Schematic of the structural transformation of a 4-membered ring of SiO<sub>4</sub> tetrahedra (right) to a smaller ring of 4 Si atoms (yellow) connected to a central C atom (dark blue) by replacing 2 O atoms (red) in the former with a C atoms, <span class="html-italic">i.e.</span>, replacement of a SiO<sub>2</sub> units with SiC. The dashed lines denote spheres (circles in two-dimension) circumscribing the 4 Si atoms in each case. The diameter of the 4-membered ring is obtained from zeolite crystal structures containing such rings [<a href="#B28-nanomaterials-05-00366" class="html-bibr">28</a>]. On the other hand, the diameter of the sphere to the left is twice the typical length of Si–C bonds (1.9 Å) in silicon carbide crystal structure.</p>
Full article ">Figure 4
<p>A view of SiOC network obtained in a previous reverse-monte-carlo (RMC) simulation study [<a href="#B19-nanomaterials-05-00366" class="html-bibr">19</a>]. Si, O and C atoms are shown in red, green and blue, respectively. Note the inhomogeneous distribution of C atoms implying partial spatial segregation of C containing SiO<span class="html-italic"><sub>x</sub></span>C<sub>4<span class="html-italic">−x</span></sub> tetrahedra along continuous channel-like regions.</p>
Full article ">
2502 KiB  
Article
In Situ Self Assembly of Nanocomposites: Competition of Chaotic Advection and Interfacial Effects as Observed by X-Ray Diffreaction
by Dilru R. Ratnaweera, Chaitra Mahesha, David A. Zumbrunnen and Dvora Perahia
Nanomaterials 2015, 5(1), 351-365; https://doi.org/10.3390/nano5010351 - 17 Mar 2015
Cited by 2 | Viewed by 5163
Abstract
The effects of chaotic advection on the in situ assembly of a hierarchal nanocomposite of Poly Amide 6, (nylon 6 or PA6) and platelet shape nanoparticles (NPs) were studied. The assemblies were formed by chaotic advection, where melts of pristine PA6 and a [...] Read more.
The effects of chaotic advection on the in situ assembly of a hierarchal nanocomposite of Poly Amide 6, (nylon 6 or PA6) and platelet shape nanoparticles (NPs) were studied. The assemblies were formed by chaotic advection, where melts of pristine PA6 and a mixture of PA6 with NPs were segregated into discrete layers and extruded into film in a continuous process. The process assembles the nanocomposite into alternating pristine-polymer and oriented NP/polymer layers. The structure of these hierarchal assemblies was probed by X-rays as a processing parameter, N, was varied. This parameter provides a measure of the extent of in situ structuring by chaotic advection. We found that all assemblies are semi-crystalline at room temperature. Increasing N impacts the ratio of α to γ crystalline forms. The effects of the chaotic advection vary with the concentration of the NPs. For nanocomposites with lower NP concentrations the amount of the γ crystalline form increased with N. However, at higher NP concentrations, interfacial effects of the NP play a significant role in determining the structure, where the NPs oriented along the melt flow direction and the polymer chains oriented perpendicular to the NP surfaces. Full article
(This article belongs to the Special Issue Self-Assembled Nanomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Transmission electron microscopy (TEM) image of (<b>a</b>) 2 vol % nanoparticles (NPs) montmorillonite nanocomposite formed by chaotic advection at <span class="html-italic">N =</span> 9 [<a href="#B27-nanomaterials-05-00351" class="html-bibr">27</a>]; and (<b>b</b>) 5.6 vol % at <span class="html-italic">N =</span> 7, extruded as 150 micron thick films. Dark regions correspond to the NP rich areas and light regions correspond to the matrix polymer.</p>
Full article ">Figure 2
<p>Schematic presentation of the agitator in chaotic advection and the time evolution of the slow. The agitation scheme used in the current study for <span class="html-italic">N =</span> 1 is presented.</p>
Full article ">Figure 3
<p>Directions of incident X-ray beams with respect to the extrusion direction. <span class="html-italic">X</span> and <span class="html-italic">Y</span> define the plane of the film and <span class="html-italic">Z</span> corresponds to the extrusion directions.</p>
Full article ">Figure 4
<p>X-ray patterns and analysis of Poly Amide 6 (PA6) films (<b>a</b>) Pristine; and (<b>b</b>) Chaotically blended at <span class="html-italic">N =</span> 30.</p>
Full article ">Figure 5
<p>(<b>a</b>) Powder X-ray patterns of extruded PA6 with and without NPs at <span class="html-italic">N</span> = 0; and (<b>b</b>) Peak deconvolution of X-ray data for 5.6 vol % nanocomposite.</p>
Full article ">Figure 6
<p>Variation of α/γ crystalline ratios with the processing parameter <span class="html-italic">N</span> for samples at the indicated NP compositions. The solid lines are drawn as a guideline for the eye.</p>
Full article ">Figure 7
<p>Total crystalline/amorphous ratio variation with respect to the processing parameter <span class="html-italic">N</span> at the indicated NP compositions. Lines are drawn as a visual guide.</p>
Full article ">Figure 8
<p>(<b>a</b>) Two-dimensional X-ray pattern of PA6 films at <span class="html-italic">N =</span> 0 and <span class="html-italic">N =</span> 30 from edge, through and end directions. The white regions in the middle of the images correspond to the shadow of the beam stopper; (<b>b</b>) The intensity along the γ crystalline ring marked in γ for the different directions; and (<b>c</b>) The cross sections of the <span class="html-italic">N =</span> 30 along the <span class="html-italic">X</span> and <span class="html-italic">Y</span> directions.</p>
Full article ">Figure 9
<p>(<b>a</b>) Two dimensional X-ray patterns of PA6 with 5.6 vol % NP nanocomposite films at <span class="html-italic">N =</span> 0 and <span class="html-italic">N =</span> 20 from edge, through and end directions; (<b>b</b>) The intensity along the γ crystalline ring marked in γ for the different directions; and (<b>c</b>) The cross sections of the <span class="html-italic">N =</span> 0 along <span class="html-italic">X</span> and <span class="html-italic">Y</span> directions.</p>
Full article ">Figure 10
<p>A schematic representation of the relative orientations of NPs and PA6 chains in nanocomposites.</p>
Full article ">
1794 KiB  
Review
Hybrids of Nucleic Acids and Carbon Nanotubes for Nanobiotechnology
by Kazuo Umemura
Nanomaterials 2015, 5(1), 321-350; https://doi.org/10.3390/nano5010321 - 12 Mar 2015
Cited by 53 | Viewed by 7829
Abstract
Recent progress in the combination of nucleic acids and carbon nanotubes (CNTs) has been briefly reviewed here. Since discovering the hybridization phenomenon of DNA molecules and CNTs in 2003, a large amount of fundamental and applied research has been carried out. Among thousands [...] Read more.
Recent progress in the combination of nucleic acids and carbon nanotubes (CNTs) has been briefly reviewed here. Since discovering the hybridization phenomenon of DNA molecules and CNTs in 2003, a large amount of fundamental and applied research has been carried out. Among thousands of papers published since 2003, approximately 240 papers focused on biological applications were selected and categorized based on the types of nucleic acids used, but not the types of CNTs. This survey revealed that the hybridization phenomenon is strongly affected by various factors, such as DNA sequences, and for this reason, fundamental studies on the hybridization phenomenon are important. Additionally, many research groups have proposed numerous practical applications, such as nanobiosensors. The goal of this review is to provide perspective on biological applications using hybrids of nucleic acids and CNTs. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>A theoretical model of structure of single-stranded DNA (ssDNA) and single-walled carbon nanotubes (SWNTs) hybrids, proposed by Zheng <span class="html-italic">et al.</span> (10,0) SWNT and poly(T) were assumed for the calculation (Reprinted from reference [<a href="#B1-nanomaterials-05-00321" class="html-bibr">1</a>] with permission).</p>
Full article ">Figure 2
<p>Optical absorption spectra of separated SWNTs by selective adsorption of ssDNA, reported by Tu <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B91-nanomaterials-05-00321" class="html-bibr">91</a>] with permission).</p>
Full article ">Figure 3
<p>Differences in photoluminescence (PL) maps due to the pH values of solvents indicated by Kim <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B89-nanomaterials-05-00321" class="html-bibr">89</a>] with permission).</p>
Full article ">Figure 4
<p>A typical force curve during the peeling of a T<sub>100</sub> ssDNA molecule away from a SWNT surface, demonstrated by Iliafar <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B83-nanomaterials-05-00321" class="html-bibr">83</a>] with permission).</p>
Full article ">Figure 5
<p>An example of agarose gel electrophoresis of DNA-SWNT hybrids with and without ssDNA-binding (SSB) proteins as performed by Nii <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B123-nanomaterials-05-00321" class="html-bibr">123</a>] with permission).</p>
Full article ">Figure 6
<p>Recent theoretical models of several combinations of ssDNA and SWNTs, proposed by Roxbury <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B169-nanomaterials-05-00321" class="html-bibr">169</a>] with permission).</p>
Full article ">Figure 7
<p>Schematic representations and AFM images of structures fabricated with ssDNA-SWNT hybrids using the DNA origami technique, generated by Mangalum <span class="html-italic">et al.</span> (Reprinted from reference [<a href="#B193-nanomaterials-05-00321" class="html-bibr">193</a>] with permission).</p>
Full article ">
5150 KiB  
Article
Chemokine-Releasing Nanoparticles for Manipulation of the Lymph Node Microenvironment
by Taissia G. Popova, Allison Teunis, Ruben Magni, Alessandra Luchini, Virginia Espina, Lance A. Liotta and Serguei G. Popov
Nanomaterials 2015, 5(1), 298-320; https://doi.org/10.3390/nano5010298 - 5 Mar 2015
Cited by 12 | Viewed by 7206
Abstract
Chemokines (CKs) secreted by the host cells into surrounding tissue establish concentration gradients directing the migration of leukocytes. We propose an in vivo CK gradient remodeling approach based on sustained release of CKs by the crosslinked poly(N-isopropylacrylamide) hydrogel open meshwork nano-particles (NPs) containing [...] Read more.
Chemokines (CKs) secreted by the host cells into surrounding tissue establish concentration gradients directing the migration of leukocytes. We propose an in vivo CK gradient remodeling approach based on sustained release of CKs by the crosslinked poly(N-isopropylacrylamide) hydrogel open meshwork nano-particles (NPs) containing internal crosslinked dye affinity baits for a reversible CK binding and release. The sustained release is based on a new principle of affinity off-rate tuning. The NPs with Cibacron Blue F3G-A and Reactive Blue-4 baits demonstrated a low-micromolar affinity binding to IL-8, MIP-2, and MCP-1 with a half-life of several hours at 37 °C. The capacity of NPs loaded with IL-8 and MIP-1α to increase neutrophil recruitment to lymph nodes (LNs) was tested in mice after footpad injection. Fluorescently-labeled NPs used as tracers indicated the delivery into the sub-capsular compartment of draining LNs. The animals administered the CK-loaded NPs demonstrated a widening of the sub-capsular space and a strong LN influx of leukocytes, while mice injected with control NPs without CKs or bolus doses of soluble CKs alone showed only a marginal neutrophil response. This technology provides a new means to therapeutically direct or restore immune cell traffic, and can also be employed for simultaneous therapy delivery. Full article
(This article belongs to the Special Issue Nanoparticles in Theranostics)
Show Figures

Figure 1

Figure 1
<p>Binding isotherms of IL-8 (0.25 ng/mL) and MCP-1 (1 ng/mL) in PBS at 4 °C, 18 h with different concentrations of NPs (5% wet volume suspension diluted as indicated). Triangles and squares correspond to Cibacron, and diamonds to Reactive Blue. [F] and [B] stand for concentrations of free and particle-bound CKs, respectively.</p>
Full article ">Figure 2
<p>Release of IL-8 and MIP-2 from reactive Blue and Cibacron NPs. CKs (2 μg/mL) were mixed with indicated NPs (10% <span class="html-italic">v</span>/<span class="html-italic">v</span> suspension) in 100 μL of three-fold diluted PBS (1/3 PBS) and incubated at 4 °C overnight. After incubation the NPs were pelleted for 5 min at 16,000 g and room temperature, re-suspended in 1 mL of 1/3 PBS, and incubated at 22 °C (<b>A</b>) or 37 °C (<b>B</b>). Portions of the suspension were withdrawn at indicated times, the NPs were pelleted for 5 min at 16,000 <span class="html-italic">g</span>, and supernatants removed. The remaining pellets were boiled for 5 min in the SDS loading buffer and the amount of CK in solution determined by Western blot. In (<b>B</b>) the 1/3 PBS buffer was supplemented with 1 mg/mL BSA. The experiments were run in duplicate. Error bars indicate SD of relative band intensities for pairwise measurements (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 3
<p>Effect of BSA on MIP-2 binding with and release from the Cibacron NPs. The CK (2 μg/mL) was loaded onto NPs (10% suspension) at 4 °C overnight in PBS buffer diluted 1:3 and supplemented with indicated concentrations of BSA in a total volume of 100 μL. After loading the particles were pelleted and Sups removed. The NP pellet was re-suspended in 1 mL of PBS with the indicated concentrations of BSA at room temperature. The amount of bound CK was determined by Western blot as described in Materials and Methods. The blot image was quantitated and relative intensities of the bands calculated. Error bars indicate SD calculated for three independent samples of control CK loaded on the same gel.</p>
Full article ">Figure 4
<p>Fluorescent pNIPAm NPs labeled with Alexa Fluor 555 (yellow) quickly migrate to sub-capsular and medullar regions of popliteal LNs of mice (arrows). A suspension of NPs (20 μL) in PBS was injected into mouse hind foot pads for 30 min and the popliteal LNs surgically removed for histologic evaluation. The LNs were paraffin-embedded after fixation with paraformaldehyde, and the 8 μm tissue slices were mounted onto glass slide. The particles were observed at 555/570 nm using Olympus BX51 microscope with a TRITC filter set. Similar responses were detected in all three mice in the group challenged with CK-loaded NPs.</p>
Full article ">Figure 5
<p>Representative images of the subcapsular and medullary regions of popliteal LNs after injection of Reactive Blue NP suspension (5% wet <span class="html-italic">v</span>/<span class="html-italic">v</span>, 50 µL) into each of the hind footpads of mice. H&amp;E-stained sections after 30 min (<b>two top rows</b>) and 4 h (<b>two bottom rows</b>) post injection. Squared regions are shown on the right at higher magnification. Neutrophils (arrows) were immunostained brown for myeloperoxidase.</p>
Full article ">Figure 6
<p>Representative images of the subcapsular and medullary regions of popliteal LNs after injection of the soluble (<b>two left columns</b>) and the Reactive Blue NP-loaded (<b>two right columns</b>) IL-8 and MIP-1α. H&amp;E-stained sections after 4 h (<b>two top rows</b>) and 24 h (<b>two bottom rows</b>). The injected amount was 5 ng and 50 ng of each CK in the total volume of 50 µL. Neutrophils were immunostained brown for myeloperoxidase.</p>
Full article ">Figure 7
<p>Enumeration of neutrophil counts in the subcapsular (<b>A</b>) and medullary (<b>B</b>) regions of popliteal LNs in the experiments corresponding to <a href="#nanomaterials-05-00298-f005" class="html-fig">Figure 5</a> and <a href="#nanomaterials-05-00298-f006" class="html-fig">Figure 6</a>. Mice were injected with the soluble and the Reactive Blue NP-loaded IL-8 and MIP-1α. The number of myeloperoxidase-positive neutrophils in the tissues slices was counted under microscope in five randomly selected fields of view (0.02 mm<sup>2</sup> each) for each of the indicated conditions (chemokine dosage, presence of NPs, and time post challenge). Error bars correspond to 95% confidence intervals. * and # indicate <span class="html-italic">p</span> ≤ 0.05 between the corresponding counts with and without NPs.</p>
Full article ">Figure 8
<p>Chemical structures of the Trypan Blue, Cibacron Blue F3G-A (Cibacron), and Reactive Blue 4 (Reactive Blue) dyes.</p>
Full article ">Figure 9
<p>ELISA of IL-8 binding (250 pg/mL) to NPs containing different baits. The particle suspension (5% wet <span class="html-italic">v</span>/<span class="html-italic">v</span>) of the NPs was incubated with IL-8 for 30 min at room temperature. Diamonds correspond to the standard IL-8 concentrations used to draw a linear calibration plot.</p>
Full article ">Figure 10
<p>Binding of MIP-2 (200 ng/mL) with Reactive Blue NP suspension (10% wet <span class="html-italic">v</span>/<span class="html-italic">v</span>) in PBS and three-fold diluted PBS at 4 °C overnight. After incubation, the particles were quickly pelleted and the amount of CK determined by Western blot. Upper panels, the Western blot images; lower panels, band intensities corresponding to the control amount of CK before binding to the NPs (C), NP pellet (P), and Sups (S).</p>
Full article ">Figure 11
<p>Raw data illustrating western blots in the experiments presented in <a href="#nanomaterials-05-00298-f002" class="html-fig">Figure 2</a> and <a href="#nanomaterials-05-00298-f003" class="html-fig">Figure 3</a>. (<b>A</b>) IL-8 release from Cibacron and Reactive Blue NPs, no BSA, 22 °C; (<b>B</b>) MIP-2, Cibacron NPs, no BSA, 22 °C; (<b>C</b>) MIP-2 release from Cibacron NPs, no BSA, 37 °C; (<b>D</b>) IL-8 release from Cibacron NPs, no BSA, 37 °C; (<b>E</b>) MIP-2 release from Cibacron NPs, with BSA, 22 °C. Numbers in the legends indicate time (h) after initiation of dissociation.</p>
Full article ">Figure 11 Cont.
<p>Raw data illustrating western blots in the experiments presented in <a href="#nanomaterials-05-00298-f002" class="html-fig">Figure 2</a> and <a href="#nanomaterials-05-00298-f003" class="html-fig">Figure 3</a>. (<b>A</b>) IL-8 release from Cibacron and Reactive Blue NPs, no BSA, 22 °C; (<b>B</b>) MIP-2, Cibacron NPs, no BSA, 22 °C; (<b>C</b>) MIP-2 release from Cibacron NPs, no BSA, 37 °C; (<b>D</b>) IL-8 release from Cibacron NPs, no BSA, 37 °C; (<b>E</b>) MIP-2 release from Cibacron NPs, with BSA, 22 °C. Numbers in the legends indicate time (h) after initiation of dissociation.</p>
Full article ">
1349 KiB  
Article
DNA/Ag Nanoparticles as Antibacterial Agents against Gram-Negative Bacteria
by Tomomi Takeshima, Yuya Tada, Norihito Sakaguchi, Fumio Watari and Bunshi Fugetsu
Nanomaterials 2015, 5(1), 284-297; https://doi.org/10.3390/nano5010284 - 3 Mar 2015
Cited by 34 | Viewed by 9488
Abstract
Silver (Ag) nanoparticles were produced using DNA extracted from salmon milt as templates. Particles spherical in shape with an average diameter smaller than 10 nm were obtained. The nanoparticles consisted of Ag as the core with an outermost thin layer of DNA. The [...] Read more.
Silver (Ag) nanoparticles were produced using DNA extracted from salmon milt as templates. Particles spherical in shape with an average diameter smaller than 10 nm were obtained. The nanoparticles consisted of Ag as the core with an outermost thin layer of DNA. The DNA/Ag hybrid nanoparticles were immobilized over the surface of cotton based fabrics and their antibacterial efficiency was evaluated using E. coli as the typical Gram-negative bacteria. The antibacterial experiments were performed according to the Antibacterial Standard of Japanese Association for the Functional Evaluation of Textiles. The fabrics modified with DNA/Ag nanoparticles showed a high enough inhibitory and killing efficiency against E. coli at a concentration of Ag ≥ 10 ppm. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>High-resolution field emission transmission electron microscope (FE-TEM) images of the DNA/Ag nanoparticles. The insets in (<b>A</b>) and (<b>B</b>) show the corresponding selected area electron diffraction pattern (a1) recorded from the whole area (<b>A</b>) and the nano-beam diffraction pattern (b1) from the area marked with the dashed line circle (b), respectively. (<b>C</b>) and (<b>D</b>) are the enlarged images from the areas indicated with the dotted square (c) in (<b>B</b>) and with the solid square (d) in (<b>C</b>), respectively.</p>
Full article ">Figure 2
<p>X-ray powder diffraction (XRD) patterns of the DNA/Ag nanoparticles.</p>
Full article ">Figure 3
<p>Number of colonies after incubation with a DNA/Ag nanoparticle aqueous suspension at different concentrations. (<b>A</b>) 1.04 × 10<sup>6</sup> colony-forming units (CFU)/mL; (<b>B</b>) 1.04 × 10<sup>7</sup> CFU/mL, of <span class="html-italic">E. coli</span>; (<b>C</b>) 8.65 × 10<sup>5</sup> CFU/mL; (<b>D</b>) 8.65 × 10<sup>6</sup> CFU/mL, of <span class="html-italic">S. aureus</span> were applied (without DNA/Ag nanoparticles, <span class="html-italic">n</span> = 2; with DNA/Ag nanoparticles, <span class="html-italic">n</span> = 4).</p>
Full article ">Figure 4
<p>(<b>A</b>) Schematic description of the procedure for immobilizing DNA/Ag nanoparticles on to the cotton fabric; (<b>B</b>) Photographs of the DNA/Ag nanoparticle immobilized fabric samples. The amounts of immobilized Ag nanoparticles are shown as Ag (ppm).</p>
Full article ">Figure 5
<p>Ag(I) ions release profiles for the DNA/Ag nanoparticle immobilized cotton fabric (34,000 ppm as Ag). Error bars represent maximum and minimum value (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 6
<p>Numbers of colonies found for the untreated fabrics (controls) and for the DNA/Ag nanoparticles immobilized fabrics (10 and 30 ppm as Ag) at the starting point (0 h) and after 18 h incubation (18 h). If no colonies were observed, the number was regarded as less than 20 CFU/mL and was shown as 20 CFU/mL Error bars represent maximum and minimum value (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>Proposed formation mechanism of the DNA/Ag nanoparticles.</p>
Full article ">
1192 KiB  
Article
DNA-Assisted Solubilization of Carbon Nanotubes and Construction of DNA-MWCNT Cross-Linked Hybrid Hydrogels
by Anatoly Zinchenko, Yosuke Taki, Vladimir G. Sergeyev and Shizuaki Murata
Nanomaterials 2015, 5(1), 270-283; https://doi.org/10.3390/nano5010270 - 3 Mar 2015
Cited by 24 | Viewed by 7319
Abstract
A simple method for preparation of DNA-carbon nanotubes hybrid hydrogel based on a two-step procedure including: (i) solubilization of multi-walled carbon nanotubes (MWCNT) in aqueous solution of DNA, and (ii) chemical cross-linking between solubilized MWCNT via adsorbed DNA and free DNA by ethylene [...] Read more.
A simple method for preparation of DNA-carbon nanotubes hybrid hydrogel based on a two-step procedure including: (i) solubilization of multi-walled carbon nanotubes (MWCNT) in aqueous solution of DNA, and (ii) chemical cross-linking between solubilized MWCNT via adsorbed DNA and free DNA by ethylene glycol diglycidyl ether is reported. We show that there exists a critical concentration of MWCNT below which a homogeneous dispersion of MWCNT in hybrid hydrogel can be achieved, while at higher concentrations of MWCNT the aggregation of MWCNT inside hydrogel occurs. The strengthening effect of carbon nanotube in the process of hydrogel shrinking in solutions with high salt concentration was demonstrated and significant passivation of MWCNT adsorption properties towards low-molecular-weight aromatic binders due to DNA adsorption on MWCNT surface was revealed. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Photographic images of MWCNT (1 mg) after dispersion in pure water (control) and in solution of 0.33% (<span class="html-italic">w</span>/<span class="html-italic">w</span>) DNA (300 bp) by sonication at 10 W for 2.5 h; (<b>B</b>) UV-Vis absorbance spectra of MWCNT aqueous solutions after dispersion by sonication in water (line <b>1</b>) and in 0.33% DNA solution (line <b>2</b>).</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>) Typical transmission electron microscopy images of MWCNT dispersed in solution of DNA. (<b>C</b>) Dynamic light scattering analysis data of MWCNT size dispersed in solution of DNA. (<b>D</b>) Zeta potential analysis data of a charge on MWCNT dispersed in solution of DNA.</p>
Full article ">Figure 3
<p>Photographic images of DNA and DNA–MWCNT hybrid hydrogel films prepared at different concentrations of MWCNT in solution (<b>A</b>: 0 mg/L (DNA hydrogel without nanotubes); <b>B</b>: 14 mg/L; <b>C</b>: 28 mg/L; <b>D</b>: 55 mg/L; <b>E</b>: 110 mg/L) after swelling in 1 mM NaCl solution. White arrows on image E indicate the flocks of aggregated nanotubes.</p>
Full article ">Figure 4
<p>Relative decrease of DNA and DNA–MWCNT hybrid hydrogel film size in solutions of varied NaCl concentrations.</p>
Full article ">Figure 5
<p>Time-dependent uptake of 1-naphtylamine from 7 mL of 20 μM solution by bare MWCNT (<b>A</b>); dispersion of carbon nanotubes in DNA solutions (<b>B</b>); DNA–MWCNT hydrogel (<b>C</b>); and DNA hydrogel (<b>D</b>). The amount of MWCNT in every sample was 1 mg.</p>
Full article ">Figure 6
<p>Schematic illustration of DNA–multi-walled carbon nanotubes (MWCNT) hybrid hydrogel preparation procedure. (<b>A</b>) Bundles of carbon nanotubes; (<b>B</b>) Individual nanotubes dispersed in solution due to solubilization by DNA under sonication; and (<b>C</b>) DNA-modified nanotubes cross-linked with an additionally added free DNA (red) using EGDE cross-linking agent at alkaline pH and elevated temperatures.</p>
Full article ">
0 pages, 231 KiB  
Retraction
RETRACTED: Begum et al. Potential Impact of Multi-Walled Carbon Nanotubes Exposure to the Seedling Stage of Selected Plant Species. Nanomaterials 2014, 4, 203–221
by Nanomaterials Editorial Office
Nanomaterials 2015, 5(1), 268-269; https://doi.org/10.3390/nano5010268 - 2 Mar 2015
Viewed by 5618
Abstract
We have become aware that a substantial part of the main text of [1] is copied from multiple other publications. In total, 46% of the main text was taken from publications by the same authors [2,3] and 10% from other papers [4,5]. Because [...] Read more.
We have become aware that a substantial part of the main text of [1] is copied from multiple other publications. In total, 46% of the main text was taken from publications by the same authors [2,3] and 10% from other papers [4,5]. Because of the extent of text taken verbatim from previously published articles, we have made the decision to retract the article. All the authors of [1] have agreed to this decision. This paper is thus declared retracted and shall be marked accordingly for the scientific record.[...] Full article
1633 KiB  
Review
DNA under Force: Mechanics, Electrostatics, and Hydration
by Jingqiang Li, Sithara S. Wijeratne, Xiangyun Qiu and Ching-Hwa Kiang
Nanomaterials 2015, 5(1), 246-267; https://doi.org/10.3390/nano5010246 - 25 Feb 2015
Cited by 22 | Viewed by 8098
Abstract
Quantifying the basic intra- and inter-molecular forces of DNA has helped us to better understand and further predict the behavior of DNA. Single molecule technique elucidates the mechanics of DNA under applied external forces, sometimes under extreme forces. On the other hand, ensemble [...] Read more.
Quantifying the basic intra- and inter-molecular forces of DNA has helped us to better understand and further predict the behavior of DNA. Single molecule technique elucidates the mechanics of DNA under applied external forces, sometimes under extreme forces. On the other hand, ensemble studies of DNA molecular force allow us to extend our understanding of DNA molecules under other forces such as electrostatic and hydration forces. Using a variety of techniques, we can have a comprehensive understanding of DNA molecular forces, which is crucial in unraveling the complex DNA functions in living cells as well as in designing a system that utilizes the unique properties of DNA in nanotechnology. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of a single molecule manipulation experiment using atomic force microscopy (AFM). (<b>a</b>) One end of the DNA molecule is attached to a substrate while the other end is pulled by the AFM cantilever tip; (<b>b</b>) The cantilever spring obeys Hooke’s law and the elasticity of DNA follows the wormlike chain (WLC) model. The stage position, λ, is related to the molecular end-to-end distance, <span class="html-italic">z</span>, by <span class="html-italic">z</span> = λ – Δ<span class="html-italic">z</span>. Adapted from [<a href="#B33-nanomaterials-05-00246" class="html-bibr">33</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Velocity dependence of the rupture force. Data showing the most probable rupture force as a function of stretching velocity for the short dsDNA of 30 bp, 20 bp, and 10 bp. The rupture force scales linearly with the logarithm of velocity. Adapted from [<a href="#B48-nanomaterials-05-00246" class="html-bibr">48</a>]. Copyright (1999) National Academy of Sciences, USA; (<b>b</b>) Dependence of the rupture force on the length of dsDNA. The central horizontal line is the critical force of the overstretching transition measured in the λ-phage DNA. The upper and lower lines correspond to 10% and 90% of overstretching transition, respectively. Adapted from [<a href="#B39-nanomaterials-05-00246" class="html-bibr">39</a>]. Reprinted with permission. Copyright (2005) American Physical Society.</p>
Full article ">Figure 3
<p>Experimental force-extension data for the stretching of λ-phage DNA. (<b>a</b>) The data are obtained from optical tweezers and fitted with the WLC model. Adapted from [<a href="#B1-nanomaterials-05-00246" class="html-bibr">1</a>]. Reprinted with permission from AAAS; (<b>b</b>) Data are obtained from AFM single molecule experiments and fitted with different one-dimensional polymer models. Adapted from [<a href="#B33-nanomaterials-05-00246" class="html-bibr">33</a>]. Reprinted with permission from Springer Science and Business Media.</p>
Full article ">Figure 4
<p>Force-extension curves for ssDNA. (<b>a</b>) Comparison of pulling curves between poly(dT) and poly(dA). Reprinted with permission from [<a href="#B60-nanomaterials-05-00246" class="html-bibr">60</a>]. Copyright (2007) by the American Physical Society; (<b>b</b>) Force-extension curves for different forms of dsDNA and ssDNA. Reprinted with permission from [<a href="#B3-nanomaterials-05-00246" class="html-bibr">3</a>]. Copyright (2010) by the American Physical Society.</p>
Full article ">Figure 5
<p>A cartoon illustration of DNA–DNA interactions in a side-by-side configuration.</p>
Full article ">Figure 6
<p>Visualization of Debye–Hückel potentials between two charged spheres of 3 nm diameter and 10 e bare charge. The effect of salt screening is shown for the case of 10 mM monovalent salt.</p>
Full article ">Figure 7
<p>Application of small angle X-ray scattering (SAXS) to measure the structure factor <span class="html-italic">S</span>(<span class="html-italic">Q</span>) of semi-dilute dispersions of oligomeric DNAs. The top panels show the case of strong DNA–DNA repulsion, giving rise to spatial ordering of DNA strands and a structure factor with pronounced correlation peaks. The bottom panels show the case of weak DNA–DNA interaction, giving rise to random dispersions and a much suppressed structure factor.</p>
Full article ">Figure 8
<p>Effective charges (z<sub>eff</sub>) determined by SAXS measurements in conjunction with physical modeling. The oligomeric DNA has a bare charge of 48 e. Symbols are experimental values and lines are the renormalized charges. Details of experimental conditions and modeling procedures are described in [<a href="#B91-nanomaterials-05-00246" class="html-bibr">91</a>]. Adapted from [<a href="#B91-nanomaterials-05-00246" class="html-bibr">91</a>]. Copyright (2006) by the American Physical Society.</p>
Full article ">Figure 9
<p>Condensed dsDNA helices packaged in hexagonal arrays, as viewed from the axis. Reproduced from [<a href="#B102-nanomaterials-05-00246" class="html-bibr">102</a>]. Copyright (2013) with permission from Elsevier.</p>
Full article ">Figure 10
<p>(<b>a</b>) Illustration of DNA arrays under osmotic stress; (<b>b</b>) Demonstration of shifts of the DNA–DNA X-ray diffraction peak as the osmotic pressure increases under DNA-condensing conditions. The peaks are scaled to assist visualization; (<b>c</b>) The force-spacing curves of dsDNA in varied salts as annotated in the legend. The <span class="html-italic">x</span>-axis shows the inter-axial spacing, whereas the surface-to-surface spacing is &lt;2 nm. Symbols are experimental data and lines are the fits using exponential forces, as described in the text. The arrow indicates the DNA spacings at zero osmotic pressure.</p>
Full article ">
726 KiB  
Review
New Insights into Understanding Irreversible and Reversible Lithium Storage within SiOC and SiCN Ceramics
by Magdalena Graczyk-Zajac, Lukas Mirko Reinold, Jan Kaspar, Pradeep Vallachira Warriam Sasikumar, Gian-Domenico Soraru and Ralf Riedel
Nanomaterials 2015, 5(1), 233-245; https://doi.org/10.3390/nano5010233 - 24 Feb 2015
Cited by 53 | Viewed by 8174
Abstract
Within this work we define structural properties of the silicon carbonitride (SiCN) and silicon oxycarbide (SiOC) ceramics which determine the reversible and irreversible lithium storage capacities, long cycling stability and define the major differences in the lithium storage in SiCN and SiOC. For [...] Read more.
Within this work we define structural properties of the silicon carbonitride (SiCN) and silicon oxycarbide (SiOC) ceramics which determine the reversible and irreversible lithium storage capacities, long cycling stability and define the major differences in the lithium storage in SiCN and SiOC. For both ceramics, we correlate the first cycle lithiation or delithiation capacity and cycling stability with the amount of SiCN/SiOC matrix or free carbon phase, respectively. The first cycle lithiation and delithiation capacities of SiOC materials do not depend on the amount of free carbon, while for SiCN the capacity increases with the amount of carbon to reach a threshold value at ~50% of carbon phase. Replacing oxygen with nitrogen renders the mixed bond Si-tetrahedra unable to sequester lithium. Lithium is more attracted by oxygen in the SiOC network due to the more ionic character of Si-O bonds. This brings about very high initial lithiation capacities, even at low carbon content. If oxygen is replaced by nitrogen, the ceramic network becomes less attractive for lithium ions due to the more covalent character of Si-N bonds and lower electron density on the nitrogen atom. This explains the significant difference in electrochemical behavior which is observed for carbon-poor SiCN and SiOC materials. Full article
Show Figures

Figure 1

Figure 1
<p>Dependence of the insertion and extraction capacity of SiOC-derived materials on the amount of free carbon (<b>a</b>) and SiOC matrix (<b>b</b>). Cycling stability defined as the ratio of the extraction capacity after prolonged cycling (&lt;100 cycles) to the first extraction capacity. Experimental data for samples pyrolysed at 1000–1100 °C from [<a href="#B54-nanomaterials-05-00233" class="html-bibr">54</a>,<a href="#B56-nanomaterials-05-00233" class="html-bibr">56</a>,<a href="#B57-nanomaterials-05-00233" class="html-bibr">57</a>].</p>
Full article ">Figure 2
<p>Dependence of the insertion and extraction capacity of SiCN-derived materials on the amount of free carbon (<b>a</b>) and SiOC matrix (<b>b</b>). Cycling stability defined as the ratio of the extraction capacity after prolonged cycling (&gt;100 cycles) to the first extraction capacity. Experimental data for samples pyrolysed at 1000–1100 °C from [<a href="#B49-nanomaterials-05-00233" class="html-bibr">49</a>,<a href="#B51-nanomaterials-05-00233" class="html-bibr">51</a>,<a href="#B53-nanomaterials-05-00233" class="html-bibr">53</a>,<a href="#B56-nanomaterials-05-00233" class="html-bibr">56</a>].</p>
Full article ">Figure 3
<p>Schematic scheme of the electronic density of states for a-SiO2, SiOC and free carbon according to Reference [<a href="#B66-nanomaterials-05-00233" class="html-bibr">66</a>].</p>
Full article ">
1056 KiB  
Article
Synthesis of Upconversion β-NaYF4:Nd3+/Yb3+/Er3+ Particles with Enhanced Luminescent Intensity through Control of Morphology and Phase
by Yunfei Shang, Shuwei Hao, Jing Liu, Meiling Tan, Ning Wang, Chunhui Yang and Guanying Chen
Nanomaterials 2015, 5(1), 218-232; https://doi.org/10.3390/nano5010218 - 24 Feb 2015
Cited by 42 | Viewed by 10243
Abstract
Hexagonal NaYF4:Nd3+/Yb3+/Er3+ microcrystals and nanocrystals with well-defined morphologies and sizes have been synthesized via a hydrothermal route. The rational control of initial reaction conditions can not only result in upconversion (UC) micro and nanocrystals with varying [...] Read more.
Hexagonal NaYF4:Nd3+/Yb3+/Er3+ microcrystals and nanocrystals with well-defined morphologies and sizes have been synthesized via a hydrothermal route. The rational control of initial reaction conditions can not only result in upconversion (UC) micro and nanocrystals with varying morphologies, but also can produce enhanced and tailored upconversion emissions from the Yb3+/Er3+ ion pairs sensitized by the Nd3+ ions. The increase of reaction time converts the phase of NaYF4:Nd3+/Yb3+/Er3+ particles from the cubic to the hexagonal structure. The added amount of oleic acid plays a critical role in the shape evolution of the final products due to their preferential attachment to some crystal planes. The adjustment of the molar ratio of F/Ln3+ can range the morphologies of the β-NaYF4:Nd3+/Yb3+/Er3+ microcrystals from spheres to nanorods. When excited by 808 nm infrared laser, β-NaYF4:Nd3+/Yb3+/Er3+ microplates exhibit a much stronger UC emission intensity than particles with other morphologies. This phase- and morphology-dependent UC emission holds promise for applications in photonic devices and biological studies. Full article
(This article belongs to the Special Issue Current Trends in Up-Converting Nanoparticles)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffraction (XRD) patterns of NaYF<sub>4</sub>: 10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals synthesized at different hydrothermal times: (<b>a</b>) 3 h; (<b>b</b>) 6 h; (<b>c</b>) 12 h; and (<b>d</b>) 24 h (180 °C F<sup>−</sup>/Ln<sup>3+</sup> = 5:1 OA/Ln<sup>3+</sup> = 40:1).</p>
Full article ">Figure 2
<p>Typical field emission scanning electron microscopy (FESEM) images for the resulting NaYF<sub>4</sub>:10% Nd<sup>3+</sup>, 10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals prepared with different hydrothermal times: (<b>a</b>) 3 h; (<b>b</b>) 6 h; (<b>c</b>) 12 h; and (<b>d</b>) 24 h (Other synthetic parameters were kept identical, 180 °C, F<sup>−</sup>/Ln<sup>3+</sup> = 5:1, OA/Ln<sup>3+</sup> = 40:1).</p>
Full article ">Figure 3
<p>Typical FESEM images for the as-prepared NaYF<sub>4</sub>: 10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals at different molar ratios of OA/Ln<sup>3+</sup>: (<b>a</b>) 6:1; (<b>b</b>) 20:1; (<b>c</b>) 30:1; and (<b>d</b>) 40:1 (Other synthetic parameters were kept identical, 180 °C, 24 h, F<sup>−</sup>/Ln<sup>3+</sup> = 5:1).</p>
Full article ">Figure 4
<p>XRD patterns of NaYF<sub>4</sub>:10% Nd<sup>3+</sup>, 10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals synthesized at different molar ratios of OA/Ln<sup>3+</sup>: (<b>a</b>) 6:1; (<b>b</b>) 20:1; (<b>c</b>) 3:1; and (<b>d</b>) 40:1 (Other synthetic parameters were kept identical, 180 °C, 24 h, F<sup>−</sup>/Ln<sup>3+</sup> = 5:1).</p>
Full article ">Figure 5
<p>Typical FESEM images for the as-prepared NaYF<sub>4</sub>: 10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals at different molar ratios of F<sup>−</sup>/Ln<sup>3+</sup>: (<b>a</b>) 1:1; (<b>b</b>) 2:1; (<b>c</b>) 3:1; and (<b>d</b>) 5:1 (Other synthetic parameters were kept identical, 180 °C, 24 h, OA/Ln<sup>3+</sup> = 20:1).</p>
Full article ">Figure 5 Cont.
<p>Typical FESEM images for the as-prepared NaYF<sub>4</sub>: 10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals at different molar ratios of F<sup>−</sup>/Ln<sup>3+</sup>: (<b>a</b>) 1:1; (<b>b</b>) 2:1; (<b>c</b>) 3:1; and (<b>d</b>) 5:1 (Other synthetic parameters were kept identical, 180 °C, 24 h, OA/Ln<sup>3+</sup> = 20:1).</p>
Full article ">Figure 6
<p>XRD patterns of NaYF<sub>4</sub>:10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals synthesized at different molar ratios of F<sup>−</sup>/Ln<sup>3+</sup>: (<b>a</b>) 1:1; (<b>b</b>) 2:1; (<b>c</b>) 3:1; and (<b>d</b>) 5:1 (Other synthetic parameters were kept identical, 180 °C, 24 h, OA/Ln<sup>3+</sup> = 20:1).</p>
Full article ">Figure 7
<p>UC Photoluminescence (PL) spectra of NaYF<sub>4</sub>:10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals with different molar ratios of OA/Ln<sup>3+</sup> under diode laser excitation at 808 nm. Excitation power density, ~390 W/cm<sup>2</sup>.</p>
Full article ">Figure 8
<p>UC Photoluminescence (PL) spectra of NaYF<sub>4</sub>:10% Nd<sup>3+</sup>,10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> microcrystals with different molar ratios of F<sup>−</sup>/Ln<sup>3+</sup> (180 °C, 24 h, OA/Ln<sup>3+</sup> = 40) under diode laser excitation at 808 nm. Excitation power density, ~390 W/cm<sup>2</sup>.</p>
Full article ">Figure 9
<p>Pump power dependence of the fluorescent bands centered at 520, 540, and 655 nm from NaYF<sub>4</sub>:10% Nd<sup>3+</sup>, 10% Yb<sup>3+</sup>, 2% Er<sup>3+</sup> on pumping power.</p>
Full article ">Figure 10
<p>Proposed energy transfer mechanism of Nd<sup>3+</sup>, Yb<sup>3+</sup>, Er<sup>3+</sup> ions following diode laser excitation of 808 nm. The dashed-dotted, dashed, dotted, and full arrows represent the photon excitation, energy transfer, and emission processes.</p>
Full article ">Figure 11
<p>Schematic illustration of the possible formation mechanism of NaYF<sub>4</sub> with various morphologies under different OA/Ln<sup>3+</sup> and F<sup>−</sup>/Ln<sup>3+</sup> molar ratio conditions.</p>
Full article ">
2596 KiB  
Article
Polymorphic Ring-Shaped Molecular Clusters Made of Shape-Variable Building Blocks
by Keitel Cervantes-Salguero, Shogo Hamada, Shin-ichiro M. Nomura and Satoshi Murata
Nanomaterials 2015, 5(1), 208-217; https://doi.org/10.3390/nano5010208 - 16 Feb 2015
Cited by 7 | Viewed by 7148
Abstract
Self-assembling molecular building blocks able to dynamically change their shapes, is a concept that would offer a route to reconfigurable systems. Although simulation studies predict novel properties useful for applications in diverse fields, such kinds of building blocks, have not been implemented thus [...] Read more.
Self-assembling molecular building blocks able to dynamically change their shapes, is a concept that would offer a route to reconfigurable systems. Although simulation studies predict novel properties useful for applications in diverse fields, such kinds of building blocks, have not been implemented thus far with molecules. Here, we report shape-variable building blocks fabricated by DNA self-assembly. Blocks are movable enough to undergo shape transitions along geometrical ranges. Blocks connect to each other and assemble into polymorphic ring-shaped clusters via the stacking of DNA blunt-ends. Reconfiguration of the polymorphic clusters is achieved by the surface diffusion on mica substrate in response to a monovalent salt concentration. This work could inspire novel reconfigurable self-assembling systems for applications in molecular robotics. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Shape-variable building block and design with DNA origami. (<b>a</b>) Abstract representation of the block with two bonding arms; (<b>b</b>) 2-mers in <span class="html-italic">cis</span> configuration and assembled polymorphic clusters (from 3-mers up to 6-mers); (<b>c</b>) (<b>left</b>): Origami scaffold (black), flexible area (mustard), contacting edges (red) and blunt-ends with binary codes (sky-blue); (<b>right</b>): Detail of the flexible area showing staples in colour (yellow stars indicate scaffold loops for tuning springs); (<b>d</b>) Origami profile and its correspondent abstract representation for the 60° and 120° configurations.</p>
Full article ">Figure 2
<p>Polymorphic clusters made of shape-variable monomers before “reconfiguration protocol” (<b>a</b>–<b>d</b>) and during the “reconfiguration protocol” (<b>e</b>). (<b>a</b>–<b>b</b>) Clusters of <span class="html-italic">M</span> (11,11). First row: cluster representations. Second row: AFM images. 3-mers, 4-mers and 5-mers (<b>a</b>) and open 6-mers (<b>b</b>) at 2 nM concentration. (<b>c</b>–<b>d</b>) Distribution of monomers contributing to the formation of <span class="html-italic">x</span>-mers in open and closed states (<b>c</b>) and ratio open:closed <span class="html-italic">x</span>-mers (<b>d</b>) for <span class="html-italic">M</span> (11,9), <span class="html-italic">M</span> (11,11) and <span class="html-italic">M</span> (11,18). <span class="html-italic">U</span> indicates unclear monomers. (<b>e</b>) Inset of the frames in movie in <a href="#app1-nanomaterials-05-00208" class="html-app">Supplementary Information</a> (0.02 fps). A 3-mers closes and two 2-mers self-assemble into an open 4-mer. AFM images in (<b>a</b>) are 310 nm × 300 nm in size. Error bars in (<b>c</b>) and SE in (<b>d</b>) indicate the standard error. <span class="html-italic">n</span> indicates the number of analysed AFM images of 2040 nm × 1680 nm, and the number of counted monomers per each AFM image is shown in <a href="#app1-nanomaterials-05-00208" class="html-app">Figure S12</a>.</p>
Full article ">Figure 3
<p>Self-assemble and reconfiguration of fixed monomers. (<b>a.1</b>) and (<b>b.1</b>): Fixed monomers. (<b>a.2–3</b>) 3-mers and cluster distribution before and after reconfiguration of <span class="html-italic">M</span> (11,0). (<b>b.2–3</b>) Representative AFM images of M (0,11) before and after reconfiguration. (<b>b.4–12</b>) Clusters after reconfiguration for different spring B. White arrows in (<b>b.5</b>) show extra M13. (<b>c</b>): Distribution of monomers contributing to the formation of <span class="html-italic">x</span>-mers including open and closed states (<b>c.1</b>) and ratio open:closed <span class="html-italic">x</span>-mers (<b>c.2</b>) for M (0,11). U indicates unclear monomers. AFM images are 310 nm × 300 nm. Width of (<b>b.9</b>) is 500 nm. Scale bars are 200 nm. Error bars in (<b>a.3, c.1</b>) and SE in (<b>c.2</b>) indicate the standard error. <span class="html-italic">n</span> indicates the number of analysed AFM images of 2040 nm × 1680 nm, and the number of counted monomers per each AFM image is shown in <a href="#app1-nanomaterials-05-00208" class="html-app">Figure S12</a>.</p>
Full article ">Figure 4
<p>Distributions of the clusters after reconfiguration for <span class="html-italic">M</span> (0,<span class="html-italic">b</span>) (<span class="html-italic">b</span> = 8−18). (<b>a</b>) Distribution of monomers contributing to the formation of <span class="html-italic">x</span>-mers including open and closed states. Colours indicate each type of cluster. Yellow indicates open 7-mer. Analysed AFM areas are 2240 nm × 1680 nm; (<b>b</b>) Ratio open:closed clusters for 4-mers, 5-mers and 6-mers. Error bars in (<b>a</b>) and SE in (<b>b</b>) indicate the standard error. <span class="html-italic">n</span> indicates the number of analysed AFM images of 2040 nm × 1680 nm, and the number of counted monomers per each AFM image is shown in <a href="#app1-nanomaterials-05-00208" class="html-app">Figure S12</a>.</p>
Full article ">
4745 KiB  
Review
DNA-Protected Silver Clusters for Nanophotonics
by Elisabeth Gwinn, Danielle Schultz, Stacy M. Copp and Steven Swasey
Nanomaterials 2015, 5(1), 180-207; https://doi.org/10.3390/nano5010180 - 12 Feb 2015
Cited by 104 | Viewed by 11782
Abstract
DNA-protected silver clusters (AgN-DNA) possess unique fluorescence properties that depend on the specific DNA template that stabilizes the cluster. They exhibit peak emission wavelengths that range across the visible and near-IR spectrum. This wide color palette, combined with low toxicity, high [...] Read more.
DNA-protected silver clusters (AgN-DNA) possess unique fluorescence properties that depend on the specific DNA template that stabilizes the cluster. They exhibit peak emission wavelengths that range across the visible and near-IR spectrum. This wide color palette, combined with low toxicity, high fluorescence quantum yields of some clusters, low synthesis costs, small cluster sizes and compatibility with DNA are enabling many applications that employ AgN-DNA. Here we review what is known about the underlying composition and structure of AgN-DNA, and how these relate to the optical properties of these fascinating, hybrid biomolecule-metal cluster nanomaterials. We place AgN-DNA in the general context of ligand-stabilized metal clusters and compare their properties to those of other noble metal clusters stabilized by small molecule ligands. The methods used to isolate pure AgN-DNA for analysis of composition and for studies of solution and single-emitter optical properties are discussed. We give a brief overview of structurally sensitive chiroptical studies, both theoretical and experimental, and review experiments on bringing silver clusters of distinct size and color into nanoscale DNA assemblies. Progress towards using DNA scaffolds to assemble multi-cluster arrays is also reviewed. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental absorbance spectra are strikingly different for ligand-protected clusters composed of the same number of metal atoms but with different shapes and type of metal. (<b>a</b>) Ag<sub>N</sub> clusters protected by different DNA template strands each display a single, narrow absorbance transition beyond the 260 nm DNA peak [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>]; (<b>b</b>) A proposed Ag<sub>N</sub>-DNA structure [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>], whose neutral cluster core (gray) is held by base-bound Ag<sup>+</sup> (blue). The depicted <span class="html-italic">rod</span> cluster shape reproduces the striking spectral purity and size-dependent colors that are characteristic of Ag<sub>N</sub>-DNA; (<b>c</b>) The broad spectra of Au<sub>N</sub>-SR clusters protected by thiolate (SR) ligands reflect their <span class="html-italic">globular</span> cluster shape and Au composition. (Adapted from reference [<a href="#B15-nanomaterials-05-00180" class="html-bibr">15</a>] and reprinted with permission); (<b>d</b>) The globular core of Au<sub>N</sub>-SR is protected by SR ligands that incorporate additional Au atoms. Orange: Au. Yellow: S. (R groups not shown). (Adapted with permission from reference [<a href="#B16-nanomaterials-05-00180" class="html-bibr">16</a>]).</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic of the HPLC-mass spectrometry (MS) system used to determine composition and optical properties of pure Ag<sub>N</sub>-DNA [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>,<a href="#B28-nanomaterials-05-00180" class="html-bibr">28</a>]; (<b>b</b>) Absorbance chromatogram, A<sub>260</sub>, from the 1st stage of purification by HPLC, with TEAA/TEA buffer, shows multiple peaks corresponding to many different Ag-bearing DNA species in the as-synthesized solution. The boxed peak is for the red fluorescent Ag<sub>N</sub>-DNA; (<b>c</b>) A<sub>260</sub> chromatogram for the 2nd HPLC stage, with HFIP/TEA buffer, which separates out remaining impurities. This enables inline MS (<b>d</b>) to identify the red emitter as an Ag<sub>15</sub>-DNA containing just one DNA strand (<span class="html-italic">n<sub>s</sub></span> = 1); (<b>e</b>) We separately identify M and Z by resolving the isotope peak envelope in high resolution MS. Alignment of measured (red) and calculated isotope patterns (bars) determines the total (<span class="html-italic">N</span>), neutral (<span class="html-italic">N</span><sub>0</sub>) and cationic (<span class="html-italic">N<sub>+</sub></span>) Ag content in Ag<sub>N</sub>-DNA, where the total Ag content <span class="html-italic">N = N</span><sub>0</sub> <span class="html-italic">+ N<sub>+</sub></span>. (Adapted with permission from references [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>,<a href="#B28-nanomaterials-05-00180" class="html-bibr">28</a>]).</p>
Full article ">Figure 3
<p>Identification of the composition of a weakly-bound strand dimer (<span class="html-italic">n<sub>s</sub></span> = 2) Ag<sub>N</sub>-DNA. Chromatograms are from the 2nd stage of HPLC purification (HFIP/TEA) with inline MS. Reprinted with permission from reference [<a href="#B28-nanomaterials-05-00180" class="html-bibr">28</a>]. (<b>a</b>) A<sub>260</sub> peaks for elution of all DNA-containing species. The fluorescent Ag<sub>N</sub>-DNA elutes at 19.5 minutes; (<b>b</b>,<b>c</b>) MS and emission spectra corresponding to the 19.5 min A<sub>260</sub> peak in (<b>a</b>); (<b>b</b>) The <span class="html-italic">Z</span> = −4, −5, −6 and −7 charge ladder (blue) identifies <span class="html-italic">N</span> = 21 as the total silver content and <span class="html-italic">n<sub>s</sub></span> = 2 as the number of DNA strands. The additional strand monomer peaks correspond to <span class="html-italic">n<sub>s</sub> =</span> 1 and <span class="html-italic">N</span> = 10 (purple) and 11 (green) fragments (as expected, the integrated fragment counts are equal within experimental error); (<b>d</b>) Fluorescence chromatogram at 558 nm; and (<b>e</b>) Mass chromatogram of the <span class="html-italic">N</span><sub>Ag</sub> = 21, <span class="html-italic">n<sub>s</sub></span> = 2 IR emitter, for <span class="html-italic">M</span>/<span class="html-italic">Z</span> = 1868.</p>
Full article ">Figure 4
<p>Sensitivity of cluster absorbance spectra to the cluster shape, size and composition. (<b>a</b>) An atomic silver rod, with the 2 types of collective excitations illustrated by their electron density profiles (color). Valence electrons slosh longitudinal to the rod in the “L” plasmon, and transverse to the rod in the “T” plasmon; (<b>b</b>) These L and T collective excitations dominate the optical absorbance spectra of Ag rods. Much weaker transitions are single particle-like. (Adapted with permission from reference [<a href="#B50-nanomaterials-05-00180" class="html-bibr">50</a>]); (<b>c</b>) When the same number of Ag atoms is rearranged from a <span class="html-italic">rod</span> (red) into a <span class="html-italic">globule</span> (green), the absorbance shifts to higher energy and becomes more scattered. (Adapted with permission from reference [<a href="#B51-nanomaterials-05-00180" class="html-bibr">51</a>]); (<b>d</b>) Calculated absorbance spectra of Ag<sub>12</sub> (top) and Au<sub>12</sub> (bottom) <span class="html-italic">rod</span> clusters. The congested spectrum for Au<sub>12</sub> arises from mixing of valence electron and <span class="html-italic">d</span> orbital transitions. (Adapted with permission from reference [<a href="#B53-nanomaterials-05-00180" class="html-bibr">53</a>]).</p>
Full article ">Figure 5
<p>Optical properties of pure Ag<sub>N</sub>-DNA. (Adapted with permission from references [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>,<a href="#B55-nanomaterials-05-00180" class="html-bibr">55</a>]). (<b>a</b>) Fluorescent Ag<sub>N</sub>-DNA show a single absorbance peak (blue) at energies below the DNA peak. Black: DNA alone; (<b>b</b>) Peak absorbance energies of fluorescent Ag<sub>N</sub>-DNA (red) follow the same trend as predicted for Ag cluster rods with 1-atom cross section (green curve) [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>]. Blue shading: Range of main absorbance peaks for same-size globular clusters. Gray curve: predictions for a thicker Ag cluster rod [<a href="#B53-nanomaterials-05-00180" class="html-bibr">53</a>]; (<b>c</b>) Another example rod structure that is consistent with optical data and the measured <span class="html-italic">N</span><sub>0</sub> and <span class="html-italic">N</span><sub>+</sub>. The Ag° rod (gray) is shown attached to DNA bases via Ag<sup>+</sup> (blue) [<a href="#B14-nanomaterials-05-00180" class="html-bibr">14</a>]; (<b>d</b>–<b>g</b>) Single Ag<sub>15</sub>-DNA at 2K (excited at 590 nm) [<a href="#B55-nanomaterials-05-00180" class="html-bibr">55</a>]; (<b>d</b>) Wide field image shows bright spots from individual Ag<sub>15</sub>-DNA; (<b>e</b>) Observation of 1-step blinking between bright and dark states confirms emission from a single Ag<sub>15</sub>-DNA; (<b>f</b>) Spectra for individual Ag<sub>15</sub>-DNA remain spectrally broad at 2K, consistent with collective transitions [<a href="#B55-nanomaterials-05-00180" class="html-bibr">55</a>]; (<b>g</b>) Dependence of emission on polarization for 2 individual Ag<sub>15</sub>-DNA. The high modulation index is expected for rod clusters [<a href="#B58-nanomaterials-05-00180" class="html-bibr">58</a>].</p>
Full article ">Figure 6
<p>Magic colors and magic number neutral cluster sizes in (<b>a</b>,<b>b</b>) Ag<sub>N</sub>-DNA and chiroptical properties of pure (<b>c</b>,<b>d</b>) Ag<sub>N</sub>-DNA. (Adapted with permission from references [<a href="#B26-nanomaterials-05-00180" class="html-bibr">26</a>,<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>]). (<b>a</b>) Fluorescence colors measured for Ag<sub>N</sub>-DNA formed on many different DNA templates exhibit “magic colors”: high green abundances near 540 nm and high red abundances near 630 nm [<a href="#B26-nanomaterials-05-00180" class="html-bibr">26</a>]; (<b>b</b>) <span class="html-italic">N</span><sub>0</sub> = 4 and 6 are magic (highly abundant) cluster sizes across all Ag<sub>N</sub>-DNA with composition determined by ESI-MS [<a href="#B26-nanomaterials-05-00180" class="html-bibr">26</a>]. Dashed lines: spherical “superatom” magic numbers 2 and 8 are <b><span class="html-italic">not</span></b> magic for Ag<sub>N</sub>-DNA, indicating non-spherical (rod) shapes for the silver clusters; (<b>c</b>) Calculated circular dichroism spectra for bare, chiral Ag cluster rods show a consistent pattern of positive and negative peaks for different curvatures [<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>]; (<b>d</b>) CD data on pure Ag<sub>N</sub>-DNA show the same peak pattern as theory [<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>].</p>
Full article ">Figure 7
<p>(Adapted with permission from [<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>,<a href="#B63-nanomaterials-05-00180" class="html-bibr">63</a>,<a href="#B64-nanomaterials-05-00180" class="html-bibr">64</a>]). (<b>a</b>,<b>b</b>) <span class="html-italic">Hybridization</span>-tuned equilibrium between dark and fluorescent forms of pure Ag<sub>N</sub>-DNA with 10–12 Ag atoms per DNA assembly. (<b>a</b>) Varying temperature, <span class="html-italic">T</span>, varies hybridization extent, producing an absorbance isosbestic point (~460 nm), that indicates an equilibrium between isomers. Peaks are at 400 nm (dark form) and 490 nm (fluorescent form) [<a href="#B63-nanomaterials-05-00180" class="html-bibr">63</a>]; (<b>b</b>) Emission (red) and absorbance at intermediate T, with both cluster forms present. Excitation at 490 nm produces emission near 570 nm, but excitation near 400 nm produces no emission [<a href="#B63-nanomaterials-05-00180" class="html-bibr">63</a>]; (<b>c</b>,<b>d</b>) <span class="html-italic">Solvent</span>-tuned equilibrium between dark and fluorescent forms of pure Ag<sub>10</sub>-DNA [<a href="#B41-nanomaterials-05-00180" class="html-bibr">41</a>]; (<b>c</b>) Absorbance for MeOH fractions of 0% to 50% of the solvent, in 5% increments. The isosbestic point (near 450 nm) indicates equilibrium between two forms of the Ag<sub>10</sub>-DNA. Absorbance peaks at 400 nm (dark form) and 490 nm (fluorescent form) [<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>], just as for the hybridization-tuned equilibrium in (<b>a</b>,<b>b</b>); (<b>d</b>) Fluorescence excitation spectra (dotted) and emission spectra at 0% and 50% MeOH. The increase in intensity is due to the higher fraction of the fluorescent form [<a href="#B42-nanomaterials-05-00180" class="html-bibr">42</a>]; (<b>e</b>,<b>f</b>) Studies of Ag-DNA and Ag-Cu-DNA made by reducing Cu<sup>2+</sup> on the fluorescent Ag-DNA solution [<a href="#B64-nanomaterials-05-00180" class="html-bibr">64</a>]; (<b>e</b>) Absorbance of unpurified solutions shows two peaks at similar wavelengths for Ag-DNA (1) and Ag-Cu-DNA (2), close to those for the dark and fluorescent cluster forms of pure Ag<sub>N</sub>-DNA in (<b>a</b>) and (<b>c</b>). The poorly resolved peak structure in the unpurified solutions in (<b>e</b>) indicate the presence of additional cluster species. (<b>f</b>) Emission from the unpurified Ag-DNA solution (1) increased upon reduction of Cu<sup>2+</sup> (2). The 560 nm emission lies near the 560–570 nm emission peaks in the pure Ag<sub>N</sub>-DNAs in (<b>a</b>–<b>d</b>), suggesting that Cu<sup>2+</sup> may also be controlling an equilibrium between dark and fluorescent cluster forms.</p>
Full article ">Figure 8
<p>Methods used to establish that multi-base motifs are key to formation of fluorescent Ag<sub>N</sub>-DNA, to recognize discriminative base motifs within DNA templates and to construct new templates for solutions with increased brightness. (Adapted with permission from reference [<a href="#B70-nanomaterials-05-00180" class="html-bibr">70</a>]). (<b>a</b>) Robotic synthesis on nearly 700 randomly generated 10-base DNA templates, and fluorescence measurement one week later, created a large, unbiased data set without unstable fluorescent products; (<b>b</b>) By running support vector machine (SVM) classifiers with different choices of feature vector, we established that multibase motifs are key to formation of fluorescent Ag<sub>N</sub>-DNA. A motif-miner algorithm, MERCI, was optimized to identify these discriminative motifs; (<b>c</b>) Combining the discovery of certain multi-base motifs important for determining fluorescence brightness with a simple generative algorithm, the probability of selecting DNA templates that stabilize fluorescent silver clusters was increased by a factor of &gt;3 relative to random templates.</p>
Full article ">Figure 9
<p>Development of dual cluster, bicolor silver cluster assemblies. Schematic to the left: Cartoons: paired cluster assemblies. Red: Ag<sub>15</sub>; Green: Ag<sub>10</sub>. (<b>a</b>) Contour map of emission intensity from the purified solution of paired clusters shows the expected peaks for direct excitation of Ag<sub>10</sub> and Ag<sub>15</sub>, and FRET: emission from Ag<sub>15</sub> due to excitation of Ag<sub>10</sub>, via radiationless energy transfer [<a href="#B82-nanomaterials-05-00180" class="html-bibr">82</a>]; (<b>b</b>) Emission excited at the Ag<sub>10</sub> absorbance peak. Green and (baseline) red curves are emission from separate solutions of the individual clusters. Orange: FRET of paired clusters produces emission from Ag<sub>15</sub> and reduces Ag<sub>10</sub> emission. The unshifted wavelengths relative to the individual clusters verify intact, stable assembly. (Adapted with permission from reference [<a href="#B82-nanomaterials-05-00180" class="html-bibr">82</a>]).</p>
Full article ">Figure 10
<p>Decoration of a DNA nanotube with fluorescent Ag<sub>N</sub>-DNA at ~1 µm separations. (Adapted with permission from reference [<a href="#B83-nanomaterials-05-00180" class="html-bibr">83</a>]). (<b>a</b>) Nanotube design, with a native DNA hairpin extruding from every tile; (<b>b</b>) Fluorescence microscopy of nanotubes assembled with hairpins on 0.5% of tiles. Direct synthesis onto the assembled tubes produces bright emission spots from fluorescent red silver clusters; (<b>c</b>) Photobleaching of emission spots along the linear nanotube contour enables estimation of ~45% initial occupation of hairpin protrusions by the unstable red clusters.</p>
Full article ">
1725 KiB  
Article
Cysteine-Functionalized Chitosan Magnetic Nano-Based Particles for the Recovery of Light and Heavy Rare Earth Metals: Uptake Kinetics and Sorption Isotherms
by Ahmed A. Galhoum, Mohammad G. Mafhouz, Sayed T. Abdel-Rehem, Nabawia A. Gomaa, Asem A. Atia, Thierry Vincent and Eric Guibal
Nanomaterials 2015, 5(1), 154-179; https://doi.org/10.3390/nano5010154 - 4 Feb 2015
Cited by 108 | Viewed by 8312
Abstract
Cysteine-functionalized chitosan magnetic nano-based particles were synthesized for the sorption of light and heavy rare earth (RE) metal ions (La(III), Nd(III) and Yb(III)). The structural, surface, and magnetic properties of nano-sized sorbent were investigated by elemental analysis, FTIR, XRD, TEM and VSM (vibrating [...] Read more.
Cysteine-functionalized chitosan magnetic nano-based particles were synthesized for the sorption of light and heavy rare earth (RE) metal ions (La(III), Nd(III) and Yb(III)). The structural, surface, and magnetic properties of nano-sized sorbent were investigated by elemental analysis, FTIR, XRD, TEM and VSM (vibrating sample magnetometry). Experimental data show that the pseudo second-order rate equation fits the kinetic profiles well, while sorption isotherms are described by the Langmuir model. Thermodynamic constants (ΔG°, ΔH°) demonstrate the spontaneous and endothermic nature of sorption. Yb(III) (heavy RE) was selectively sorbed while light RE metal ions La(III) and Nd(III) were concentrated/enriched in the solution. Cationic species RE(III) in aqueous solution can be adsorbed by the combination of chelating and anion-exchange mechanisms. The sorbent can be efficiently regenerated using acidified thiourea. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme for the synthesis of cysteine-functionalized chitosan magnetic nano-based particles.</p>
Full article ">Figure 2
<p>FTIR spectra of (<b>a</b>) chitosan-magnetite nanoparticles; (<b>b</b>) after cross-linking; (<b>c</b>) cross-linked chitosan magnetite with the spacer arm; and (<b>d</b>) cysteine-adsorbent.</p>
Full article ">Figure 3
<p>Powder X-ray diffraction (XRD) pattern of cysteine-sorbent nanoparticles.</p>
Full article ">Figure 4
<p>TEM micrographs (the scale bars are (<b>a</b>) 100 and (<b>b</b>) 50 nm, respectively).</p>
Full article ">Figure 5
<p>Magnetization curve of cysteine-functionalized chitosan magnetic nano-based particles.</p>
Full article ">Figure 6
<p>Effect of pH on the sorption of La(III), Nd(III) and Yb(III) ions using cysteine-functionalized chitosan magnetic nano-based particles (<span class="html-italic">C<sub>i</sub></span> = 100 mg·L<sup>−1</sup>; <span class="html-italic">T</span> = 300 K; <span class="html-italic">t</span> = 4 h; <span class="html-italic">m</span> = 0.05 g; <span class="html-italic">V</span> = 100 mL).</p>
Full article ">Figure 7
<p>Effect of contact time on the adsorption of La(III), Nd(III) and Yb(III) ions (<span class="html-italic">C<sub>i</sub></span> = 100 mg·L<sup>−1</sup>; <span class="html-italic">T</span> = 300 K; pH = 5; <span class="html-italic">m</span> = 0.05 g; <span class="html-italic">V</span> = 100 mL).</p>
Full article ">Figure 8
<p>Adsorption isotherms for La(III), Nd(III) and Yb(III) ions at different temperatures. (<span class="html-italic">t</span> = 4 h; <span class="html-italic">T</span> = 300 K; pH = 5; <span class="html-italic">m</span> = 0.05 g; <span class="html-italic">V</span> = 20 mL).</p>
Full article ">Figure 9
<p>Van’t Hoff plots of ln <span class="html-italic">K<sub>L</sub></span> against 1/<span class="html-italic">T</span>, for sorption of metal (III) ions.</p>
Full article ">
3293 KiB  
Communication
The Nucleotide Capture Region of Alpha Hemolysin: Insights into Nanopore Design for DNA Sequencing from Molecular Dynamics Simulations
by Richard M. A. Manara, Susana Tomasio and Syma Khalid
Nanomaterials 2015, 5(1), 144-153; https://doi.org/10.3390/nano5010144 - 27 Jan 2015
Cited by 8 | Viewed by 8418
Abstract
Nanopore technology for DNA sequencing is constantly being refined and improved. In strand sequencing a single strand of DNA is fed through a nanopore and subsequent fluctuations in the current are measured. A major hurdle is that the DNA is translocated through the [...] Read more.
Nanopore technology for DNA sequencing is constantly being refined and improved. In strand sequencing a single strand of DNA is fed through a nanopore and subsequent fluctuations in the current are measured. A major hurdle is that the DNA is translocated through the pore at a rate that is too fast for the current measurement systems. An alternative approach is “exonuclease sequencing”, in which an exonuclease is attached to the nanopore that is able to process the strand, cleaving off one base at a time. The bases then flow through the nanopore and the current is measured. This method has the advantage of potentially solving the translocation rate problem, as the speed is controlled by the exonuclease. Here we consider the practical details of exonuclease attachment to the protein alpha hemolysin. We employ molecular dynamics simulations to determine the ideal (a) distance from alpha-hemolysin, and (b) the orientation of the monophosphate nucleotides upon release from the exonuclease such that they will enter the protein. Our results indicate an almost linear decrease in the probability of entry into the protein with increasing distance of nucleotide release. The nucleotide orientation is less significant for entry into the protein. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>The alpha-hemolysin (αHL) protein in a 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) bilayer, with the mononucleotide positioned above the vestibule entrance, where each mononucleotide represents an individual simulation. The protein is represented by yellow ribbons and for other atoms: Carbon is shown in cyan, oxygen in red, nitrogen in blue, phosphorus in brown and hydrogen in white. The waters and ions are excluded for clarity. (<b>Inset</b>) The phosphate orientations used, termed the “up” and “down” orientations, shown top and bottom respectively.</p>
Full article ">Figure 2
<p>Plots showing the relaionship between release loction and probability of capture. The highest possible probability of entry for (<b>a</b>) the distance study and (<b>b</b>) the displacement study, e.g., these are the simulations where either capture or possible capture is observed. The lowest possible probability of entry for (<b>c</b>) the distance study and (<b>d</b>) the displacement study, e.g., these are the simulations where capture is observed. The black lines are for the phosphate up orientation and the red lines are the phosphate down orientation. Error bars were calculated by using the standard error, treating the data as a binomial distribution.</p>
Full article ">Figure 3
<p>The residues with the highest propensity to interact with the nucleotide for: failed capture (<b>a</b>) and possible capture (<b>b</b>). The blue highlighted residues corresponding to failed capture are on the edge of the vestibule entrance and on the surface of the vestibule. The red highlighted residues associated with possible capture are predominantly on the edge of the entrance to the vestibule. (<b>Inset</b>) The most frequently observed binding mode between cytosine monophosphate (CMP) and the amino acid triplet, D45, K46 and N47. Hydrogen bonds (dashed lines) are observed between the hydroxyl group of the sugar and the side-chain of D45, as well as the nucleobase and the side-chain of N47. These hydrogen bonds are stable and are present for extended periods of time (greater than 1 ns). Residues are coloured for clarity; D45 in red K46 in blue, and N47 in yellow.</p>
Full article ">Figure 4
<p>The probabilities of entry for given release points relative to the ring of K8 residues for (<b>a</b>) the phosphate down orientation and (<b>b</b>) the phosphate up orientation. Simulations performed with the phosphate in the down orientation are closer than the corresponding phosphate up simulations for the same capture probability.</p>
Full article ">
3684 KiB  
Review
Alumina Matrix Composites with Non-Oxide Nanoparticle Addition and Enhanced Functionalities
by Dušan Galusek and Dagmar Galusková
Nanomaterials 2015, 5(1), 115-143; https://doi.org/10.3390/nano5010115 - 27 Jan 2015
Cited by 31 | Viewed by 7720
Abstract
The addition of SiC or TiC nanoparticles to polycrystalline alumina matrix has long been known as an efficient way of improving the mechanical properties of alumina-based ceramics, especially strength, creep, and wear resistance. Recently, new types of nano-additives, such as carbon nanotubes (CNT), [...] Read more.
The addition of SiC or TiC nanoparticles to polycrystalline alumina matrix has long been known as an efficient way of improving the mechanical properties of alumina-based ceramics, especially strength, creep, and wear resistance. Recently, new types of nano-additives, such as carbon nanotubes (CNT), carbon nanofibers (CNF), and graphene sheets have been studied in order not only to improve the mechanical properties, but also to prepare materials with added functionalities, such as thermal and electrical conductivity. This paper provides a concise review of several types of alumina-based nanocomposites, evaluating the efficiency of various preparation methods and additives in terms of their influence on the properties of composites. Full article
Show Figures

Figure 1

Figure 1
<p>Microstructure refinement observed in Al<sub>2</sub>O<sub>3</sub>-SiC nanocomposites with increasing volume fraction of SiC nanoparticles: (<b>a</b>) 3 vol% SiC; (<b>b</b>) 8 vol% SiC. The nanocomposites were sintered without pressure at 1750 °C from a green body prepared by warm pressing of poly(allyl)carbosilane-coated alumina powder [<a href="#B34-nanomaterials-05-00115" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>Creep deformation of Al<sub>2</sub>O<sub>3</sub>-SiC (AS) nanocomposites measured at 1350 °C and mechanical load of 75, 150, and 200 MPa. The increase of load is reflected as a break at the stress-strain curve. The number in the sample denomination represents the volume fraction of SiC in the material, <span class="html-italic">i.e.</span>, AS5c represents the Al<sub>2</sub>O<sub>3</sub>-SiC nanocomposite with 5 vol% of SiC [<a href="#B88-nanomaterials-05-00115" class="html-bibr">88</a>].</p>
Full article ">Figure 3
<p>Composition dependence of (<b>a</b>) hardness; (<b>b</b>) fracture toughness; and (<b>c</b>) fracture strength of Al<sub>2</sub>O<sub>3</sub>-MWCNT (denoted AC) and Al<sub>2</sub>O<sub>3</sub>-ZrO<sub>2</sub>-MWCNT (denoted AZC) nanocomposites. Comparison to monolithic alumina reference [<a href="#B95-nanomaterials-05-00115" class="html-bibr">95</a>]. Red circles represent the respective properties of nanocomposite with various volume fractions of MWCNT. Black squares represent the same property of the monolithic alumina reference.</p>
Full article ">Figure 4
<p>Indication of toughening mechanisms observed in Al<sub>2</sub>O<sub>3</sub>-ZrO<sub>2</sub>-MWCNT nanocomposites [<a href="#B95-nanomaterials-05-00115" class="html-bibr">95</a>].</p>
Full article ">Figure 5
<p>Temperature dependence of thermal conductivity of Al<sub>2</sub>O<sub>3</sub>-SiC (AS) nanocomposites with various volume fractions of SiC. Comparison to monolithic alumina reference [<a href="#B126-nanomaterials-05-00115" class="html-bibr">126</a>]. The number in the sample denomination represents the volume fraction of SiC in the material (<span class="html-italic">i.e.</span>, AS3c represents the Al<sub>2</sub>O<sub>3</sub>-SiC nanocomposite containing 3 vol% SiC).</p>
Full article ">Figure 6
<p>Composition dependence of DC electric conductivity of Al<sub>2</sub>O<sub>3</sub>-SiC nanocomposites with various volume fractions of SiC. Comparison to monolithic alumina reference [<a href="#B126-nanomaterials-05-00115" class="html-bibr">126</a>].</p>
Full article ">Figure 7
<p>Composition dependence of DC electric conductivity of Al<sub>2</sub>O<sub>3</sub>-MWCNT (material AC) and Al<sub>2</sub>O<sub>3</sub>-ZrO<sub>2</sub>-MWCNT (material AZC) nanocomposites [<a href="#B95-nanomaterials-05-00115" class="html-bibr">95</a>].</p>
Full article ">Figure 8
<p>Composition dependence of thermal conductivity of Al<sub>2</sub>O<sub>3</sub>-MWCNT (material AC) and Al<sub>2</sub>O<sub>3</sub>-ZrO<sub>2</sub>-MWCNT (material AZC) nanocomposites [<a href="#B95-nanomaterials-05-00115" class="html-bibr">95</a>].</p>
Full article ">
3248 KiB  
Review
Recent Advances on Carbon Nanotubes and Graphene Reinforced Ceramics Nanocomposites
by Iftikhar Ahmad, Bahareh Yazdani and Yanqiu Zhu
Nanomaterials 2015, 5(1), 90-114; https://doi.org/10.3390/nano5010090 - 20 Jan 2015
Cited by 140 | Viewed by 11653
Abstract
Ceramics suffer the curse of extreme brittleness and demand new design philosophies and novel concepts of manufacturing to overcome such intrinsic drawbacks, in order to take advantage of most of their excellent properties. This has been one of the foremost challenges for ceramic [...] Read more.
Ceramics suffer the curse of extreme brittleness and demand new design philosophies and novel concepts of manufacturing to overcome such intrinsic drawbacks, in order to take advantage of most of their excellent properties. This has been one of the foremost challenges for ceramic material experts. Tailoring the ceramics structures at nanometre level has been a leading research frontier; whilst upgrading via reinforcing ceramic matrices with nanomaterials including the latest carbon nanotubes (CNTs) and graphene has now become an eminent practice for advanced applications. Most recently, several new strategies have indeed improved the properties of the ceramics/CNT nanocomposites, such as by tuning with dopants, new dispersions routes and modified sintering methods. The utilisation of graphene in ceramic nanocomposites, either as a solo reinforcement or as a hybrid with CNTs, is the newest development. This article will summarise the recent advances, key difficulties and potential applications of the ceramics nanocomposites reinforced with CNTs and graphene. Full article
Show Figures

Figure 1

Figure 1
<p>Statistical analysis of the carbon nanotubes (CNTs)-reinforced ceramic nanocomposites.</p>
Full article ">Figure 2
<p>Structural features of (<b>a</b>) Monolithic Al<sub>2</sub>O<sub>3</sub> showing large grains with inter-granular fracture; (<b>b</b>) CNTs/Al<sub>2</sub>O<sub>3</sub> nanocomposites with fine grains; (<b>c</b>) Trans-granular fracture mode in CNTs/Al<sub>2</sub>O<sub>3</sub> nanocomposites; and (<b>d</b>) Single-walled (SW)CNTs at grain boundary of Al<sub>2</sub>O<sub>3</sub> matrix. TEM images exhibiting the CNT–ceramic interactions (<b>e</b>) Multi-walled (MW)CNTs (black arrow) showing their morphology in nanocomposite; (<b>f</b>) A single MWCNT existing at grain boundary; (<b>g</b>) in porosity and (<b>h</b>) Embedded within a single ceramic grain. Adapted from References [<a href="#B32-nanomaterials-05-00090" class="html-bibr">32</a>] and [<a href="#B66-nanomaterials-05-00090" class="html-bibr">66</a>] with permissions. Copyright 2010, Elsevier Ltd.</p>
Full article ">Figure 3
<p>(<b>a</b>) TEM image of the pristine MWCNTs; (<b>b</b>) High-magnification TEM image of the acid-treated MWCNT surface, arrow indicates nano-pit; (<b>c</b>) Nano-pit on the acid-treated MWCNTs is filled up with Al<sub>2</sub>O<sub>3</sub> crystal; and (<b>d</b>) Rough surface of MWCNT produced by chemical vapour deposition (CVD) method. Adapted from References [<a href="#B12-nanomaterials-05-00090" class="html-bibr">12</a>] and [<a href="#B32-nanomaterials-05-00090" class="html-bibr">32</a>] with permissions. Copyright 2009, Elsevier Ltd. and 2008 IOP Publishing Ltd.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) High-resolution TEM images showing CNT/ceramic interfaces. Adapted from References [<a href="#B15-nanomaterials-05-00090" class="html-bibr">15</a>] and [<a href="#B32-nanomaterials-05-00090" class="html-bibr">32</a>] with permissions. Copyright 2005 Advanced Study Center Co. Ltd. and 2010 Elsevier Ltd.</p>
Full article ">Figure 5
<p>SEM images from fractured surface of GNT-Al<sub>2</sub>O<sub>3</sub> nanocomposites with various GNP/CNT ratio, (<b>a</b> and <b>b</b>); Al<sub>2</sub>O<sub>3</sub>-(0.5 wt% GNP + 1 wt% CNT), (<b>c–e</b>); Al<sub>2</sub>O<sub>3</sub>-(0.5 wt% GNP + 0.5 wt% CNT), (<b>f</b>); Al<sub>2</sub>O<sub>3</sub>-0.5 wt% GNP. Adapted from Reference [<a href="#B26-nanomaterials-05-00090" class="html-bibr">26</a>] with permission. Copyright 2014 Elsevier Ltd.</p>
Full article ">
309 KiB  
Review
Magnetic Properties of Magnetic Nanoparticles for Efficient Hyperthermia
by Ihab M. Obaidat, Bashar Issa and Yousef Haik
Nanomaterials 2015, 5(1), 63-89; https://doi.org/10.3390/nano5010063 - 9 Jan 2015
Cited by 382 | Viewed by 16847
Abstract
Localized magnetic hyperthermia using magnetic nanoparticles (MNPs) under the application of small magnetic fields is a promising tool for treating small or deep-seated tumors. For this method to be applicable, the amount of MNPs used should be minimized. Hence, it is essential to [...] Read more.
Localized magnetic hyperthermia using magnetic nanoparticles (MNPs) under the application of small magnetic fields is a promising tool for treating small or deep-seated tumors. For this method to be applicable, the amount of MNPs used should be minimized. Hence, it is essential to enhance the power dissipation or heating efficiency of MNPs. Several factors influence the heating efficiency of MNPs, such as the amplitude and frequency of the applied magnetic field and the structural and magnetic properties of MNPs. We discuss some of the physics principles for effective heating of MNPs focusing on the role of surface anisotropy, interface exchange anisotropy and dipolar interactions. Basic magnetic properties of MNPs such as their superparamagnetic behavior, are briefly reviewed. The influence of temperature on anisotropy and magnetization of MNPs is discussed. Recent development in self-regulated hyperthermia is briefly discussed. Some physical and practical limitations of using MNPs in magnetic hyperthermia are also briefly discussed. Full article
124 KiB  
Editorial
Acknowledgement to Reviewers of Nanomaterials in 2014
by Nanomaterials Editorial Office
Nanomaterials 2015, 5(1), 61-62; https://doi.org/10.3390/nano5010061 - 7 Jan 2015
Viewed by 3712
Abstract
The editors of Nanomaterials would like to express their sincere gratitude to the following reviewers for assessing manuscripts in 2014:[...] Full article
1604 KiB  
Article
Cellular Uptake of Tile-Assembled DNA Nanotubes
by Samet Kocabey, Hanna Meinl, Iain S. MacPherson, Valentina Cassinelli, Antonio Manetto, Simon Rothenfusser, Tim Liedl and Felix S. Lichtenegger
Nanomaterials 2015, 5(1), 47-60; https://doi.org/10.3390/nano5010047 - 30 Dec 2014
Cited by 52 | Viewed by 10085
Abstract
DNA-based nanostructures have received great attention as molecular vehicles for cellular delivery of biomolecules and cancer drugs. Here, we report on the cellular uptake of tubule-like DNA tile-assembled nanostructures 27 nm in length and 8 nm in diameter that carry siRNA molecules, folic [...] Read more.
DNA-based nanostructures have received great attention as molecular vehicles for cellular delivery of biomolecules and cancer drugs. Here, we report on the cellular uptake of tubule-like DNA tile-assembled nanostructures 27 nm in length and 8 nm in diameter that carry siRNA molecules, folic acid and fluorescent dyes. In our observations, the DNA structures are delivered to the endosome and do not reach the cytosol of the GFP-expressing HeLa cells that were used in the experiments. Consistent with this observation, no elevated silencing of the GFP gene could be detected. Furthermore, the presence of up to six molecules of folic acid on the carrier surface did not alter the uptake behavior and gene silencing. We further observed several challenges that have to be considered when performing in vitro and in vivo experiments with DNA structures: (i) DNA tile tubes consisting of 42 nt-long oligonucleotides and carrying single- or double-stranded extensions degrade within one hour in cell medium at 37 °C, while the same tubes without extensions are stable for up to eight hours. The degradation is caused mainly by the low concentration of divalent ions in the media. The lifetime in cell medium can be increased drastically by employing DNA tiles that are 84 nt long. (ii) Dyes may get cleaved from the oligonucleotides and then accumulate inside the cell close to the mitochondria, which can lead to misinterpretation of data generated by flow cytometry and fluorescence microscopy. (iii) Single-stranded DNA carrying fluorescent dyes are internalized at similar levels as the DNA tile-assembled tubes used here. Full article
(This article belongs to the Special Issue Frontiers in Nucleic Acid Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Characterization of nanotubes. (<b>a</b>) Gel electrophoresis analysis of assembled nanotubes: (1) 1-kb ladder, (2) nanotube, (3) nanotube + folate, (4) nanotube + folate + siRNA, and (5) individual oligonucleotide. Electron micrographs of (<b>b</b>) Nanotubes; (<b>c</b>) Nanotubes with folate; and (<b>d</b>) Nanotubes with folate and siRNA (scale bars: 50 nm; insets: 20 nm).</p>
Full article ">Figure 2
<p>Endosomal uptake of nanotubes in HeLa cells. Endosomal staining of nanotubes with dextran. (<b>a</b>) Nanotubes; (<b>b</b>) Dextran; (<b>c</b>) Merged image from (<b>a</b>), (<b>b</b>) and a third channel (DAPI, blue). (<b>d</b>) Flow cytometry analysis of folate-dependent uptake of Atto488-labeled nanotubes over 24 h. Untreated cells act as the control, and the specific fluorescence intensity (SFI) of the dye is depicted. (<b>e</b>) Fluorescence intensity of stably GFP-expressing Hela cells upon the addition of nanotubes carrying <span class="html-italic">GFP</span>-targeting siRNAs or upon transfection of a <span class="html-italic">GFP</span>-targeting siRNA and a non-targeting siRNA, respectively, as controls using lipofection (LF). The median fluorescence intensity (MFI) of GFP is depicted.</p>
Full article ">Figure 3
<p>Stability of nanotubes. (<b>a</b>) Stability of nanotubes in PBS with different Mg<sup>2+</sup> concentrations; (<b>b</b>) Stability of nanotubes carrying siRNA in PBS with different Mg<sup>2+</sup> concentrations; (<b>c</b>) Stability of nanotubes in DMEM medium in the absence or presence of FCS; (<b>d</b>) Stability of nanotubes carrying siRNA in DMEM medium in the absence or presence of FCS (L: 1 kb ladder; C: control. All samples were incubated at 37 °C).</p>
Full article ">Figure 4
<p>Stability of nanotubes assembled from 84 nt-long oligonucleotides. (<b>a</b>) Schematic depiction of a section of the 6HT demonstrating the hybridization of 84mers; (<b>b</b>) Schematic depiction of a section of the 6HT demonstrating the hybridization of 42mers; (<b>c</b>) Stability of nanotubes (84mers) in PBS with different Mg<sup>2+</sup> concentrations; (<b>d</b>) Stability of nanotubes (84mers and 42mers) in DMEM + 10% FCS, DMEM and PBS. Nanotubes were incubated at 45 °C for 2 h (L: 1 kb ladder; C: control).</p>
Full article ">Figure 5
<p>Effect of dye cleavage on cellular uptake. (<b>a</b>) Endosomal staining using Alexa Fluor 488-coupled dextran (shown in green) of HeLa cells treated with oligodeoxynucleotide (ODN)-Atto647 (shown in red); (<b>b</b>) Mitochondrial colocalization of Atto647 (shown in red) in HeLa cells stained with the mitochondrial dye MitoTracker Green (shown in green). Nuclei are stained with Hoechst 33342; (<b>c</b>) Flow cytometry analysis of fluorescence intensity of cells treated with ODN-Atto647 in the absence or presence of FCS; (<b>d</b>) Flow cytometry analysis of fluorescence intensity of cells treated with Atto488-dUTP, ODN-Atto488 and nanotube labeled with Atto488. Untreated cells act as the control, and the specific fluorescence intensity (SFI) of the dye is depicted.</p>
Full article ">Figure 6
<p>DNA nanotube assembly. (<b>Left</b>) Click reaction of alkyne-modified oligonucleotides with azide-modified PEGylated folate. (<b>Right</b>) Self-assembly of 24 oligonucleotides into a six-helix tube after a 17-h annealing process.</p>
Full article ">
1088 KiB  
Article
A Novel Arch-Shape Nanogenerator Based on Piezoelectric and Triboelectric Mechanism for Mechanical Energy Harvesting
by Chenyang Xue, Junyang Li, Qiang Zhang, Zhibo Zhang, Zhenyin Hai, Libo Gao, Ruiting Feng, Jun Tang, Jun Liu, Wendong Zhang and Dong Sun
Nanomaterials 2015, 5(1), 36-46; https://doi.org/10.3390/nano5010036 - 26 Dec 2014
Cited by 49 | Viewed by 10726
Abstract
A simple and cost-effective approach was developed to fabricate piezoelectric and triboelectric nanogenerator (P-TENG) with high electrical output. Additionally, pyramid micro structures fabricated atop a polydimethylsiloxane (PDMS) surface were employed to enhance the device performance. Furthermore, piezoelectric barium titanate (BT) nanoparticles and multiwalled [...] Read more.
A simple and cost-effective approach was developed to fabricate piezoelectric and triboelectric nanogenerator (P-TENG) with high electrical output. Additionally, pyramid micro structures fabricated atop a polydimethylsiloxane (PDMS) surface were employed to enhance the device performance. Furthermore, piezoelectric barium titanate (BT) nanoparticles and multiwalled carbon nanotube (MWCNT) were mixed in the PDMS film during the forming process. Meanwhile, the composition of the film was optimized to achieve output performance, and favorable toughness was achieved after thermal curing. An arch-shape ITO/PET electrode was attached to the upper side of the polarized composite film and an aluminum film was placed under it as the bottom electrode. With periodic external force at 20 Hz, electrical output of this P-TENG, reached a peak voltage of 22 V and current of 9 μA with a peak current density of 1.13 μA/cm2, which was six times that of the triboelectric generator without BT and MWCNT nanoparticles. The nanogenerator can be directly used to lighten 28 commercial light-emitting diodes (LEDs) without any energy storage unit or rectification circuit under human footfalls. Full article
(This article belongs to the Special Issue Nanomaterials for Energy and Sustainability Applications)
Show Figures

Figure 1

Figure 1
<p>Output performance of the arch-shape piezoelectric and triboelectric nanogenerator (P-TENG) with nanostructured polydimethylsiloxane (PDMS) film under external forces at different frequencies.</p>
Full article ">Figure 2
<p>(<b>a</b>) Open-circuit voltage of the arch-shaped P-TENG with different <span class="html-italic">N</span> at 20 Hz. (<b>b</b>) Open-circuit voltage of the P-TENG with different compositions based on <span class="html-italic">N</span> = 2:3, including: A. generator based on piezoelectric effect; B. generator based on triboelectric effect; C. Generator based on both the piezoelectric and triboelectric effect; (<b>c</b>) Open-circuit voltage and short-circuit current of the arch-shape nanogenerator with different compositions, when the doping radio was fixed (<span class="html-italic">N</span> = 2:3, <span class="html-italic">V</span> = 9%), including: sample A: flat PDMS film; sample B: PDMS film with only micro pyramid arrays; sample C: PDMS film with micro pyramid arrays, unpolarized barium titanate (BT) nanoparticles and carbon nanotube (CNT); sample D: PDMS with only polarized BT nanoparticles and CNT; sample E: PDMS film with micro pyramid arrays, polarized BT nanoparticles and CNT (P-TENG).</p>
Full article ">Figure 3
<p>Working principle of this arch-shape nanogenerator. (<b>a</b>–<b>e</b>) Schematic diagram shows the working principle of the arch-shape P-TENG.</p>
Full article ">Figure 4
<p>Applications of the arch-shape P-TENG. (<b>a</b>) Diagram of the P-TENG between two plexiglasses with springs; (<b>b</b>) The output open circuit voltage is about 30 V under footstep. (<b>c</b>,<b>d</b>) When footstep falls on the P-TENG, 28 paralleled commercial LEDs were lightened without using any energy storage device or rectification circuit. Notes: All LEDs are connected in serial.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) Schematic view of the structural design and fabrication process flowchart of the P-TENG device (the total size of the P-BM composite films is 2 cm × 4 cm × 0.3 mm).</p>
Full article ">Figure 6
<p>Structure characterization of the P-BM composite films. (<b>a</b>) SEM image of the BaTiO<sub>3</sub> NPs; (<b>b</b>) TEM image of multiwalled carbon nanotube (MWCNTs) with a diameter of 20 nm and a length of 20 μm; (<b>c</b>) SEM image of the pyramid PDMS thin film with BaTiO<sub>3</sub> NPs and MWCNTs; (<b>d</b>) Enlarge photograph of the PDMS thin film tore locally by tweezers; (<b>e</b>) High-magnification SEM image of the patterned PDMS surfaces with pyramids features; (<b>f</b>) The high-magnification SEM image of mixing effect about BaTiO<sub>3</sub> NPs and MWCNTs.</p>
Full article ">
688 KiB  
Article
Synthesis, Characterization, and Mechanism of Formation of Janus-Like Nanoparticles of Tantalum Silicide-Silicon (TaSi2/Si)
by Andrey V. Nomoev, Sergey P. Bardakhanov, Makoto Schreiber, Dashima Zh. Bazarova, Boris B. Baldanov and Nikolai A. Romanov
Nanomaterials 2015, 5(1), 26-35; https://doi.org/10.3390/nano5010026 - 25 Dec 2014
Cited by 15 | Viewed by 7679
Abstract
Metal-semiconductor Janus-like nanoparticles with the composition tantalum silicide-silicon (TaSi2/Si) were synthesized for the first time by means of an evaporation method utilizing a high-power electron beam. The composition of the synthesized particles were characterized using high-resolution transmission electron microscopy (HRTEM), X-ray [...] Read more.
Metal-semiconductor Janus-like nanoparticles with the composition tantalum silicide-silicon (TaSi2/Si) were synthesized for the first time by means of an evaporation method utilizing a high-power electron beam. The composition of the synthesized particles were characterized using high-resolution transmission electron microscopy (HRTEM), X-ray diffraction (XRD), selective area electron diffraction (SAED), and energy dispersive X-ray fluorescence (EDX) analysis. The system is compared to previously synthesized core-shell type particles in order to show possible differences responsible for the Janus-like structure forming instead of a core-shell architecture. It is proposed that the production of Janus-like as opposed to core-shell or monophase particles occurs due to the ability of Ta and Si to form compounds and the relative content of Ta and Si atoms in the produced vapour. Based on the results, a potential mechanism of formation for the TaSi2/Si nanoparticles is discussed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffractogram (XRD) of the obtained powder.</p>
Full article ">Figure 2
<p>Transmission electron microscopy (TEM) image of the obtained TaSi<sub>2</sub>/Si nanoparticles.</p>
Full article ">Figure 3
<p>Energy dispersive X-ray fluorescence (EDX) spectra of (<b>a</b>) The darker (TaSi<sub>2</sub>) region; and (<b>b</b>) The lighter (Si) region of the obtained TaSi<sub>2</sub>/Si JL nanoparticles.</p>
Full article ">Figure 4
<p>Representative selective area electron diffraction (SAED) diffraction patterns obtained from (<b>a</b>) The darker region (TaSi<sub>2</sub>); and (<b>b</b>) Lighter region (Si) of the obtained JL nanoparticles.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic of the Ta and Si arrangement in a graphite crucible before electron beam irradiation; (<b>b</b>) Dependence of the electron beam current on time. The materials in the crucible are melted in region (i) and evaporated in region (ii).</p>
Full article ">
1361 KiB  
Review
Rare Earth Ion-Doped Upconversion Nanocrystals: Synthesis and Surface Modification
by Hongjin Chang, Juan Xie, Baozhou Zhao, Botong Liu, Shuilin Xu, Na Ren, Xiaoji Xie, Ling Huang and Wei Huang
Nanomaterials 2015, 5(1), 1-25; https://doi.org/10.3390/nano5010001 - 25 Dec 2014
Cited by 66 | Viewed by 13999
Abstract
The unique luminescent properties exhibited by rare earth ion-doped upconversion nanocrystals (UCNPs), such as long lifetime, narrow emission line, high color purity, and high resistance to photobleaching, have made them widely used in many areas, including but not limited to high-resolution displays, new-generation [...] Read more.
The unique luminescent properties exhibited by rare earth ion-doped upconversion nanocrystals (UCNPs), such as long lifetime, narrow emission line, high color purity, and high resistance to photobleaching, have made them widely used in many areas, including but not limited to high-resolution displays, new-generation information technology, optical communication, bioimaging, and therapy. However, the inherent upconversion luminescent properties of UCNPs are influenced by various parameters, including the size, shape, crystal structure, and chemical composition of the UCNPs, and even the chosen synthesis process and the surfactant molecules used. This review will provide a complete summary on the synthesis methods and the surface modification strategies of UCNPs reported so far. Firstly, we summarize the synthesis methodologies developed in the past decades, such as thermal decomposition, thermal coprecipitation, hydro/solvothermal, sol-gel, combustion, and microwave synthesis. In the second part, five main streams of surface modification strategies for converting hydrophobic UCNPs into hydrophilic ones are elaborated. Finally, we consider the likely directions of the future development and challenges of the synthesis and surface modification, such as the large-scale production and actual applications, stability, and so on, of the UCNPs. Full article
(This article belongs to the Special Issue Current Trends in Up-Converting Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) TEM image of edge-to-edge super lattices of LaF<sub>3</sub> triangular nanoplates. Reproduced from [<a href="#B18-nanomaterials-05-00001" class="html-bibr">18</a>]. Copyright 2005, American Chemical Society. (<b>b,c</b>) TEM images of edge-to-edge and face-to-face super lattices of SmF<sub>3</sub> hexagonal nanoplates, respectively. Reproduced from [<a href="#B19-nanomaterials-05-00001" class="html-bibr">19</a>]. Copyright 2006, John Wiley and Sons.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic illustration of the crystal structure evolution at varying polarities of the reaction medium. (<b>b</b>) Upconversion luminescence spectra of hexagonal-phase NaScF<sub>4</sub>:Yb/Er and NaYF<sub>4</sub>:Yb/Er nanocrystals. Reproduced from [<a href="#B35-nanomaterials-05-00001" class="html-bibr">35</a>]. Copyright 2012, American Chemical Society.</p>
Full article ">Figure 3
<p>TEM images of (<b>a</b>) La(OH)<sub>3</sub> nanowires. Reproduced from [<a href="#B47-nanomaterials-05-00001" class="html-bibr">47</a>]. Copyright 2002, John Wiley and Sons; (<b>b</b>) Y<sub>2</sub>O<sub>3</sub> nanotubes. Reproduced from [<a href="#B48-nanomaterials-05-00001" class="html-bibr">48</a>]. Copyright 2003, John Wiley and Sons; (<b>c</b>) LaF<sub>3</sub> nanoparticles. Reproduced from [<a href="#B49-nanomaterials-05-00001" class="html-bibr">49</a>]. Copyright 2003, John Wiley and Sons; (<b>d</b>) La(OH)<sub>3</sub> nanobelt; (<b>e</b>) a typical XRD pattern of the as-synthesized La(OH)<sub>3</sub>. Reproduced from [<a href="#B50-nanomaterials-05-00001" class="html-bibr">50</a>]. Copyright 2007, John Wiley and Sons.</p>
Full article ">Figure 4
<p>TEM and SEM images of NaYF<sub>4</sub>:Yb/Er nanocrystals prepared under different hydro/solvothermal conditions. (<b>a,b</b>) TEM images of NaYF<sub>4</sub>:Yb/Er nanocrystals synthesized in acetic acid and ethanol in the presence of CTAB, respectively; (<b>c,d</b>) TEM images of NaYF<sub>4</sub>:Yb/Er nanocrystals using EDTA in acetic acid and ethanol, respectively. Reproduced from [<a href="#B34-nanomaterials-05-00001" class="html-bibr">34</a>]. Copyright 2005, John Wiley and Sons; SEM images of (<b>e,f</b>) flower-patterned hexagonal disks; (<b>g</b>) hexagonal nanotubes; and (<b>h</b>) nanorods of β-NaYF<sub>4</sub>. Reproduced from [<a href="#B51-nanomaterials-05-00001" class="html-bibr">51</a>]. Copyright 2007, John Wiley and Sons.</p>
Full article ">Figure 5
<p>TEM images of the 1.0 mol% Er<sup>3+</sup>-doped BaTiO<sub>3</sub> nanoparticles obtained after heating to three different temperatures: (<b>a</b>) 700 °C; (<b>b</b>) 850 °C; and (<b>c</b>) 1000 °C. Reproduced from [<a href="#B71-nanomaterials-05-00001" class="html-bibr">71</a>]. Copyright 2003, American Chemical Society.</p>
Full article ">Figure 6
<p>TEM images of (<b>a</b>) YF<sub>3</sub> nanophosphors prepared in the NP-5 stabilized microemulsion system. Reproduced from [<a href="#B74-nanomaterials-05-00001" class="html-bibr">74</a>]. Copyright 2005, American Chemical Society. (<b>b</b>) YF<sub>3</sub> nanobundles synthesized in the CTAB and NP-5 stabilized microemulsion system. Reproduced from [<a href="#B75-nanomaterials-05-00001" class="html-bibr">75</a>]. Copyright 2008, American Chemical Society.</p>
Full article ">Figure 7
<p>Schematic illustration of the active core-active shell nanoparticle architecture showing the absorption of NIR near infrared light by the Yb<sup>3+</sup>-rich shell (red) and subsequent energy transfer to the Er<sup>3+</sup>,Yb<sup>3+</sup>-doped core (green), which leads to upconverted blue, green, and red emissions. Reproduced from [<a href="#B91-nanomaterials-05-00001" class="html-bibr">91</a>]. Copyright 2009, John Wiley and Sons.</p>
Full article ">Figure 8
<p>Silica-coated NaYF<sub>4</sub>:Yb/Er nanocrystals and their application for cell imaging. (<b>a</b>1–C<b>a</b>3) TEM images of silica-coated NaYF<sub>4</sub>:Yb/Er UCNPs upconversion nanoparticles at different magnifications; (<b>b</b>) Confocal fluorescence image of MCF-7 cells using silica-coated NaYF<sub>4</sub>:Yb/Er nanospheres (Left: bright-field, middle: upconversion image under 980 nm excitation, and Right: superimposed images of MCF-7 cells incubated with the nanoparticles for 24 h); (<b>c</b>) Confocal fluorescence images of MCF-7 cells with the nanospheres, excited by a 980 nm laser with different power intensities. Reproduced from [<a href="#B103-nanomaterials-05-00001" class="html-bibr">103</a>]. Copyright 2008, John Wiley and Sons.</p>
Full article ">Figure 9
<p>Schematic illustration of the UCNP-based drug delivery system. (<b>a</b>) As-synthesized oleic acid capped UCNPs; (<b>b</b>) C18PMH-PEG-FA functionalized UCNPs; (<b>c</b>) DOX loading on UCNPs. DOX molecules are physically adsorbed into the oleic acid layer on the nanoparticle surface by hydrophobic interactions; (<b>d</b>) Release of DOX from UCNPs triggered by decreasing pH. Reproduced from [<a href="#B110-nanomaterials-05-00001" class="html-bibr">110</a>]. Copyright 2011, Elsevier Ltd.</p>
Full article ">Figure 10
<p>(<b>a</b>) Schematic illustration of the ligand oxidization process; (<b>b,c</b>) TEM images of the NaYF<sub>4</sub>:Yb/Er nanoparticles before and after ligand oxidization, respectively; (<b>d</b>) Luminescence spectra of a mixture of streptavidin-functionalized NaYF<sub>4</sub>:Yb/Er nanoparticles, capture-DNA, and reporter-DNA in the presence of different concentrations of target-DNA under continuous-wave excitation at 980 nm. The linear relationships between target-DNA concentration and the intensity ratios of (<b>e</b>) I540/I654 and (<b>f</b>) I580/I540. Reproduced from [<a href="#B112-nanomaterials-05-00001" class="html-bibr">112</a>]. Copyright 2008, American Chemical Society.</p>
Full article ">Figure 11
<p>(<b>a</b>) Schematic illustration showing the ligand-exchange reactions on OM-stabilized upconversion nanophosphors. (<b>b,c</b>) TEM images of NaYF<sub>4</sub>:Yb/Er nanophosphors prior to and after ligand exchange reactions, respectively. Reproduced from [<a href="#B114-nanomaterials-05-00001" class="html-bibr">114</a>]. Copyright 2006, John Wiley and Sons.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop