[go: up one dir, main page]

Next Issue
Volume 4, June
Previous Issue
Volume 3, December
 
 

Nanomaterials, Volume 4, Issue 1 (March 2014) – 10 articles , Pages 1-188

  • 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:
552 KiB  
Article
Magnetite Nanoparticles Induce Genotoxicity in the Lungs of Mice via Inflammatory Response
by Yukari Totsuka, Kousuke Ishino, Tatsuya Kato, Sumio Goto, Yukie Tada, Dai Nakae, Masatoshi Watanabe and Keiji Wakabayashi
Nanomaterials 2014, 4(1), 175-188; https://doi.org/10.3390/nano4010175 - 18 Mar 2014
Cited by 30 | Viewed by 6884
Abstract
Nanomaterials are useful for their characteristic properties and are commonly used in various fields. Nanosized-magnetite (MGT) is widely utilized in medicinal and industrial fields, whereas their toxicological properties are not well documented. A safety assessment is thus urgently required for MGT, and genotoxicity [...] Read more.
Nanomaterials are useful for their characteristic properties and are commonly used in various fields. Nanosized-magnetite (MGT) is widely utilized in medicinal and industrial fields, whereas their toxicological properties are not well documented. A safety assessment is thus urgently required for MGT, and genotoxicity is one of the most serious concerns. In the present study, we examined genotoxic effects of MGT using mice and revealed that DNA damage analyzed by a comet assay in the lungs of imprinting control region (ICR) mice intratracheally instilled with a single dose of 0.05 or 0.2 mg/animal of MGT was approximately two- to three-fold higher than that of vehicle-control animals. Furthermore, in gpt delta transgenic mice, gpt mutant frequency (MF) in the lungs of the group exposed to four consecutive doses of 0.2 mg MGT was significantly higher than in the control group. Mutation spectrum analysis showed that base substitutions were predominantly induced by MGT, among which G:C to A:T transition and G:C to T:A transversion were the most significant. To clarify the mechanism of mutation caused by MGT, we analyzed the formation of DNA adducts in the lungs of mice exposed to MGT. DNA was extracted from lungs of mice 3, 24, 72 and 168 h after intratracheal instillation of 0.2 mg/body of MGT, and digested enzymatically. 8-Oxo-7,8-dihydro-2′-deoxyguanosine (8-oxodG) and lipid peroxide-related DNA adducts were quantified by stable isotope dilution liquid chromatography-mass spectrometry (LC-MS/MS). Compared with vehicle control, these DNA adduct levels were significantly increased in the MGT-treated mice. In addition to oxidative stress- and inflammation related-DNA adduct formations, inflammatory cell infiltration and focal granulomatous formations were also observed in the lungs of MGT-treated mice. Based on these findings, it is suggested that inflammatory responses are probably involved in the genotoxicity induced by MGT in the lungs of mice. Full article
(This article belongs to the Special Issue Nanotoxicology)
Show Figures

Figure 1

Figure 1
<p>Crystal appearance and zeta potential of magnetite nanoparticles (MGT): (<b>a</b>) Scanning electron microscopy (SEM) micrographs of MGT obtained at <span class="html-italic">E</span>=20 kV,×300,000;and (<b>b</b>) Size distribution of MGT measured in water, 0.2 μg/mL.</p>
Full article ">Figure 2
<p>DNA damage in the lungs of imprinting control region (ICR) mice intratracheally instilled with MGT. DNA damage was measured by comet assay. Male micewere treated at a dose of 0.05 mg or 0.2 mg of particles per animal, and sacrificed 3 h after particle administration. The values represent the means of data for five animals ± SE. <b>**</b> <span class="html-italic">P</span> &lt; 0.01, by the Dunnett’s test after one-way analysis of variance <span class="html-italic">vs</span>. the corresponding vehicle control mice.</p>
Full article ">Figure 3
<p>The <span class="html-italic">gpt</span> mutation frequencies in the lungs of mice after multiple intratracheal instillations of MGT. Male mice were treated with multiple (0.05 or 0.2 mg/mouse × 4 times) doses of MGT, and mice were sacrificed eight weeks after MGT administration. The data represent the mean ± SD; <b>*</b> <span class="html-italic">P</span> &lt; 0.05 by the Student’s <span class="html-italic">t</span>-test <span class="html-italic">vs.</span> the corresponding vehicle control mice.</p>
Full article ">Figure 4
<p>Microscopic findings in the lungs of <span class="html-italic">gpt</span> delta mice intratracheally instilled with MGT.Representative histopathology of the lungs of: (<b>a</b>) a control mouse given vehicle (once a week for 4 weeks; killed at 22 weeks of age); and (<b>b</b>,<b>c</b>) a mouse given multiple doses of 0.2 mg MGT (killed at 22 weeks of age). The brown-colored material is MGT.</p>
Full article ">Figure 5
<p>Oxidative and lipid peroxide-related DNA adduct formation induced by MGT exposure in the lungs of ICR mice. DNA was extracted from the lungs 3, 24, 72 and 168 h after intratracheal instillation of 0.2 mg of MGT, and was digested enzymatically. Control samples were obtained from the lungs of mice given the vehicle for the same durations of MGT exposure. 8-Oxo-7,8-dihydro-2′-deoxyguanosine (8-oxodG) and two types of Hε-adduct were quantified by stable isotope dilution liquid chromatography-mass spectrometry (LC-MS/MS). Asterisks (* and **) indicate a significant difference (<span class="html-italic">P</span> &lt; 0.05 and <span class="html-italic">P</span> &lt; 0.01) from vehicle control (treatment with 0.05% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) Tween-80) at the same point in the Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
1011 KiB  
Article
Percolation Diffusion into Self-Assembled Mesoporous Silica Microfibres
by John Canning, George Huyang, Miles Ma, Alison Beavis, David Bishop, Kevin Cook, Andrew McDonagh, Dongqi Shi, Gang-Ding Peng and Maxwell J. Crossley
Nanomaterials 2014, 4(1), 157-174; https://doi.org/10.3390/nano4010157 - 10 Mar 2014
Cited by 26 | Viewed by 8442
Abstract
Percolation diffusion into long (11.5 cm) self-assembled, ordered mesoporous microfibres is studied using optical transmission and laser ablation inductive coupled mass spectrometry (LA-ICP-MS). Optical transmission based diffusion studies reveal rapid penetration (<5 s, D > 80 μm2∙s1) of [...] Read more.
Percolation diffusion into long (11.5 cm) self-assembled, ordered mesoporous microfibres is studied using optical transmission and laser ablation inductive coupled mass spectrometry (LA-ICP-MS). Optical transmission based diffusion studies reveal rapid penetration (<5 s, D > 80 μm2∙s1) of Rhodamine B with very little percolation of larger molecules such as zinc tetraphenylporphyrin (ZnTPP) observed under similar loading conditions. The failure of ZnTPP to enter the microfibre was confirmed, in higher resolution, using LA-ICP-MS. In the latter case, LA-ICP-MS was used to determine the diffusion of zinc acetate dihydrate, D~3 × 10−4 nm2∙s−1. The large differences between the molecules are accounted for by proposing ordered solvent and structure assisted accelerated diffusion of the Rhodamine B based on its hydrophilicity relative to the zinc compounds. The broader implications and applications for filtration, molecular sieves and a range of devices and uses are described. Full article
(This article belongs to the Special Issue Ordered Mesoporous Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>An illustration of hcp packing and the two types of interstitial regions. As an approximation a round sphere that can fit into the pores is often assumed to determine the size of a molecule that can percolate or diffuse into the structure. This does not, however, take into account irregular shapes and quantities of larger molecules, more typical of real situations. The site size distribution shown reflects that calculated from the bulk of the size distribution of the nanoparticles used in this work, measured by dynamic light scattering (DLS).</p>
Full article ">Figure 2
<p>The formulas, schematic and space filling structures, as well as dimensions of each of the three molecules used in this work. The Cl<sup>−</sup> anion of Rhodamine B is free in solution and may be displaced by the negatively charged water at a silica-water interface.</p>
Full article ">Figure 3
<p>Surface characterisation of a self-assembled silica microwire: (<b>a</b>) optical micrograph of long wires produced using gravity assisted deposition by evaporative self-assembly; (<b>b</b>,<b>c</b>) SEM local images of the wire surface reproduced from earlier work (Naqshbandi <span class="html-italic">et al.</span> [<a href="#B1-nanomaterials-04-00157" class="html-bibr">1</a>]) to illustrate the presence of a finite thickness layer, ~(250–500) nm, (<b>a</b>) and (<b>b</b>) on top of an inner core. An examination of the surface layer (<b>c</b>,<b>d</b>) reveals what appears to be a mix of fcc and hcp packing in places; (<b>e</b>) shows an AFM analysis of the &gt;11 cm wires used in this work with similar hcp packing to that of previous work [<a href="#B3-nanomaterials-04-00157" class="html-bibr">3</a>].</p>
Full article ">Figure 4
<p>Schematic of the optical transmission measurement through the microfibre with a drop of Rhodamine B containing water. The coupling area between standard fibre and slab microfiber is zoomed in for clarity.</p>
Full article ">Figure 5
<p>Optical transmission of 632 nm through a self-assembled wire as a function of time during percolation of Rhodamine B using the setup illustrated in <a href="#nanomaterials-04-00157-f004" class="html-fig">Figure 4</a>. Inset shows a close-up of the initial attenuation the signal has been normalized to the initial signal level measured.</p>
Full article ">Figure 6
<p>Laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS) of self-assembled wires doped with ZnTPP and zinc acetate solutions (<span class="html-italic">c</span><sub>0</sub> = 100 mM). (<b>a</b>) Signal intensity as a function of each ablation over time for both molecules and (<b>b</b>) Diffusion analysis (natural log of signal intensity <span class="html-italic">vs.</span> area) of the hydrated Zn<sup>2+</sup> from the zinc acetate solution.</p>
Full article ">
127 KiB  
Editorial
Acknowledgement to Reviewers of Nanomaterials in 2013
by Nanomaterials Editorial Office
Nanomaterials 2014, 4(1), 155-156; https://doi.org/10.3390/nano4010155 - 27 Feb 2014
Viewed by 3361
Abstract
The editors of Nanomaterials would like to express their sincere gratitude to the following reviewers for assessing manuscripts in 2013. [...] Full article
2979 KiB  
Review
Recent Advance of Biological Molecular Imaging Based on Lanthanide-Doped Upconversion-Luminescent Nanomaterials
by Yuanzeng Min, Jinming Li, Fang Liu, Parasuraman Padmanabhan, Edwin K. L. Yeow and Bengang Xing
Nanomaterials 2014, 4(1), 129-154; https://doi.org/10.3390/nano4010129 - 6 Feb 2014
Cited by 102 | Viewed by 14771
Abstract
Lanthanide-doped upconversion-luminescent nanoparticles (UCNPs), which can be excited by near-infrared (NIR) laser irradiation to emit multiplex light, have been proven to be very useful for in vitro and in vivo molecular imaging studies. In comparison with the conventionally used down-conversion fluorescence imaging strategies, [...] Read more.
Lanthanide-doped upconversion-luminescent nanoparticles (UCNPs), which can be excited by near-infrared (NIR) laser irradiation to emit multiplex light, have been proven to be very useful for in vitro and in vivo molecular imaging studies. In comparison with the conventionally used down-conversion fluorescence imaging strategies, the NIR light excited luminescence of UCNPs displays high photostability, low cytotoxicity, little background auto-fluorescence, which allows for deep tissue penetration, making them attractive as contrast agents for biomedical imaging applications. In this review, we will mainly focus on the latest development of a new type of lanthanide-doped UCNP material and its main applications for in vitro and in vivo molecular imaging and we will also discuss the challenges and future perspectives. 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>The structure and optical properties of upconversion nanoparticles (UCNPs): (<b>a</b>) Scheme illustration of the structure and components of the UCNPs; (<b>b</b>–<b>e</b>) Upconversion multicolor tuning in Ln<sup>3+</sup>-doped cubic NaYF<sub>4</sub> UCNPs. Room temperature UCNP emission spectra of (<b>b</b>) NaYF<sub>4</sub>:Yb/Er (18/2 mol%), (<b>c</b>) NaYF<sub>4</sub>:Yb/Tm (20/0.2 mol%), (<b>d</b>) NaYF<sub>4</sub>:Yb/Er (25–60/2 mol%), and (<b>e</b>) NaYF<sub>4</sub>:Yb/Tm/Er (20/0.2/0.2–1.5 mol%) particles in ethanol solutions (10 mM). The spectra in (<b>d</b>) and (<b>e</b>) were normalized to Er<sup>3+</sup> 660 nm and Tm<sup>3+</sup> 480 nm emissions, respectively; (<b>f</b>) The proposed energy transfer mechanisms showing the upconversion processes in Er<sup>3+</sup>, Tm<sup>3+</sup>, and Yb<sup>3+</sup> doped crystals under 980 nm diode laser excitation. The dashed-dotted, dashed, dotted and full arrows represent photon excitation, energy transfer, multiphoton relaxation, and emission processes, respectively. Only visible and NIR emissions are shown here. (<b>g</b>) Upconversion multicolor fine-tuning through the use of lanthanide-doped NaYF<sub>4</sub> nanoparticles with varied dopant ratios. Adapted with permission from references [<a href="#B33-nanomaterials-04-00129" class="html-bibr">33</a>,<a href="#B34-nanomaterials-04-00129" class="html-bibr">34</a>,<a href="#B35-nanomaterials-04-00129" class="html-bibr">35</a>] and [<a href="#B39-nanomaterials-04-00129" class="html-bibr">39</a>], respectively. NIR, near-infrared. Copyright: American Chemical Society, 2008; Royal Society of Chemistry, 2009; and John Wiley and Sons, 2013.</p>
Full article ">Figure 2
<p>Surface functionalization of UCNPs and their functional groups for biological applications. Adapted with permission and modified from references [<a href="#B27-nanomaterials-04-00129" class="html-bibr">27</a>,<a href="#B28-nanomaterials-04-00129" class="html-bibr">28</a>,<a href="#B29-nanomaterials-04-00129" class="html-bibr">29</a>,<a href="#B30-nanomaterials-04-00129" class="html-bibr">30</a>,<a href="#B31-nanomaterials-04-00129" class="html-bibr">31</a>]. Copyright: Royal Society of Chemistry, 2013.</p>
Full article ">Figure 3
<p><span class="html-italic">In vivo</span> imaging of living rats with quantum dots (QDs) injected into the translucent skin of the foot (<b>a</b>) showing fluorescence, but not through thicker skin of the back (<b>b</b>) or abdomen (<b>c</b>), NaYF<sub>4</sub>:Yb/Er nanoparticles injected below the abdominal skin (<b>d</b>), thigh muscles (<b>e</b>) or below the skin of the back (<b>f</b>) show luminescence. QDs on a black disk in (<b>a</b>,<b>b</b>) are used as the control. Adapted with permission from reference [<a href="#B83-nanomaterials-04-00129" class="html-bibr">83</a>]. Copyright: Elsevier, 2008.</p>
Full article ">Figure 4
<p><span class="html-italic">In vivo</span> imaging of a mouse with the injection of UCNPs: intact mouse (<b>left</b>); the same mouse after dissection (<b>right</b>). The red color indicates emission from UCNPs; green and black show the background, as indicated by the arrows. The inset presents the photoluminescence spectra corresponding to the spectrally unmixed components of the multispectral image obtained with the Maestro system. Adapted with permission from reference [<a href="#B78-nanomaterials-04-00129" class="html-bibr">78</a>]. Copyright: American Chemical Society, 2008.</p>
Full article ">Figure 5
<p>UCNPs for cellular labeling and <span class="html-italic">in vivo</span> tracking analysis. (<b>a</b>) Confocal UCNP imaging (<b>left</b>) and its overlay with a bright field image (<b>right</b>) of cells stained with 200 μg mL<sup>−1</sup> NaLuF<sub>4</sub> UCNPs for 3 h at 37 °C. (<b>b</b>) <span class="html-italic">In vivo</span> UCNPs imaging of athymic nude mice after subcutaneous injection of 50 human nasopharyngeal epidermal carcinoma KB cells (<b>left</b>) and tail-vein injection of 1000 KB cells (<b>right</b>). The KB cells were pre-incubated with 200μg mL<sup>−1</sup> NaLuF4 UCNPs for 3 h at 37 °C before injection. (<b>c</b> and <b>d</b>) <span class="html-italic">In vivo</span> detection of UCNP-labeled mMSCs (an exogenous contast agent to track mouse Mesenchymal Stem Cells). (<b>c</b>) An upconversion luminescence image of a mouse subcutaneously injected with various numbers of mouse mesenchymal stem cells (1 × 10<sup>5</sup>) labeled with UCNPs. (<b>d</b>) Quantification of UCNPs luminescence signals in (<b>c</b>). Adapted with permission from references [<a href="#B87-nanomaterials-04-00129" class="html-bibr">87</a>,<a href="#B89-nanomaterials-04-00129" class="html-bibr">89</a>], respectively. UCL, upconversion luminescence. Copyright: American Chemical Society, 2011; and Elsevier, 2012.</p>
Full article ">Figure 6
<p>(<b>a</b>) Scheme of the synthesis of UCNP-Arginine-Glycine-Asparatic (RGD). (<b>b</b>) Time-dependent <span class="html-italic">in vivo</span> upconversion luminescence imaging of subcutaneous U87 MG (left hind leg, indicated by short arrows) and MCF-7 (Michigan Cancer Foundation-7) tumor (right hind leg, indicated by long arrows) borne by athymic nude mice after intravenous injection of UCNP-RGD over a 24 h period. UCNP-RGD conjugate was prepared from UCNP-OA complex (OA: Oleylamine). All images were acquired under the same instrumental conditions (power ≈ 80 mW cm<sup>-2</sup> and temperature ≈ 21.5 °C on the surface of the mouse). Adapted with permission from reference [<a href="#B92-nanomaterials-04-00129" class="html-bibr">92</a>]. Copyright: American Chemical Society, 2009.</p>
Full article ">Figure 7
<p>The design for photo-controlled Dox delivery through mesoporous silica coated UCNPs conjugated with folic acid. Adapted with permission from reference [<a href="#B73-nanomaterials-04-00129" class="html-bibr">73</a>]. Copyright: John Wiley and Sons, 2013.</p>
Full article ">Figure 8
<p>Bioluminescent images of firefly luciferase (fLuc) activity in living mice that were treated with D-luciferin. (<b>a</b>) Experimental design for uncaging D-luciferin and subsequent bioluminescence through the use of photo-caged core-shell upconversion nanoparticles. (<b>b</b>) <b>Left</b>: injection with D-luciferin (20 μM, 20 μL); <b>right</b>: injection with photo-caged nanoparticles without NIR light irradiation. (<b>c</b>) <b>Left</b>: injection with photo-caged nanoparticles and irradiation with UV light for 10 min; <b>right</b>: injection with photo-caged nanoparticles and irradiation with NIR light for 1 h. Adapted with permission from reference [<a href="#B62-nanomaterials-04-00129" class="html-bibr">62</a>]. Copyright: John Wiley and Sons, 2012.</p>
Full article ">Figure 9
<p>Live-cell apoptosis imaging for NIR irradiation of Pt(IV)-probe UCNPs@SiO<sub>2</sub> (15 µM) incubated cells: (<b>a</b>) schematic illustration of NIR light activation of platinum(IV) prodrug and intracellular apoptosis imaging through upconversion nanoparticles; (<b>b</b>) A2780 cells; and (<b>c</b>) cisplatin resistant A2780cis cells. (blue: DAPI (4,6-Diamidino-2-phenylindole); green: Annexin V; red: Cy5.) Quantitative flow cytometric analysis of (<b>d</b>) A2780 and (<b>e</b>) A2780cis cells treated with different concentrations of Pt(IV)-probe UCNPs@SiO<sub>2</sub> (10, 15 and 20 µM, respectively) and 1 h NIR irradiation. Cells treated with Ac-DEVD (Aspartic acid-Glutamic acid-Valine-Aspartic acid)-CHO (20 µM) inhibitor and NIR irradiation of cells without Pt(IV)-probe UCNP incubation were used as controls. Adapted with permission from reference [<a href="#B102-nanomaterials-04-00129" class="html-bibr">102</a>]. Copyright: John Wiley and Sons, 2013.</p>
Full article ">Figure 10
<p><span class="html-italic">In vivo</span> Single Photon Emission Tomography (SPECT)/Optical imaging study after intravenous injection of <sup>153</sup>Sm-UCNPs. (<b>a</b>) Whole-body three-dimensional projection, (<b>b</b>) coronal, (<b>c</b>) sagittal and (<b>d</b>) transversal images acquired at 1 h and (<b>e</b>) whole-body three-dimensional projection images acquired at 24 h are shown respectively. The arrows in the inset point to the liver (L) and spleen (S). (<b>f</b>) <span class="html-italic">In vivo</span> upconversion luminescence imaging of the Kunming mouse 1 h after tail vein injection of the <sup>153</sup>Sm-UCNPs (20 mg/kg). Adapted with permission from reference [<a href="#B105-nanomaterials-04-00129" class="html-bibr">105</a>]. Copyright: Elsevier, 2013.</p>
Full article ">Figure 11
<p>(<b>I</b>) A whole-body imaging of UCNPs@SiO<sub>2</sub>-GdDTPA (Diethylenetriaminepentaacetic Acid) for 10 min. (<b>I-A</b>) <span class="html-italic">In vivo</span> imaging of the sacrificed nude mouse after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 10 min. (<b>I-B</b>) <span class="html-italic">Ex vivo</span> imaging of nude mouse. (<b>I-C</b>) <span class="html-italic">Ex vivo</span> imaging of viscera. All images were acquired under the same instrumental conditions, and the power density of the 980 nm laser is 150 mW cm<sup>−2</sup>. (<b>II</b>) The application of <span class="html-italic">in vivo</span> CT imaging in Kunming mice. (<b>II-A</b>, <b>B</b> and <b>C</b>) serials coronal CT images of Kunming mouse at different layer after injection with UCNPs@SiO<sub>2</sub>-GdDTPA. (<b>II-E</b>, <b>F</b> and <b>G</b>) partial enlarged CT view of abdomen. (<b>III</b>) The application of <span class="html-italic">in vivo</span> MRI imaging of the Kunming mice. (<b>III-A</b>) T<sub>1</sub>-weighted MR images of liver after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min (<b>III-B</b>) T<sub>1</sub> distribution images of liver after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min. (<b>III-C</b>) Local colorized T<sub>1</sub>-weighted MR images of liver after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min. (<b>III-D</b>) T<sub>1</sub>-weighted MR images of spleen after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min. (<b>III-E</b>) T<sub>1</sub> distribution images of spleen after injection with UCNPs<b>@</b>SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min. (<b>III-F</b>) Local colorized T<sub>1</sub>-weighted MR images of spleen after injection with UCNPs@SiO<sub>2</sub>-GdDTPA for 0, 30 and 120 min. Adapted with permission from reference [<a href="#B85-nanomaterials-04-00129" class="html-bibr">85</a>]. Copyright: Elsevier, 2012.</p>
Full article ">Figure 12
<p>Absorption of water in the NIR and the integration scheme of the Nd<sup>3+</sup>→Yb<sup>3+</sup> energy transfer (ET) process by introducing the Nd<sup>3+</sup>/Yb<sup>3+</sup> co-doped shell. The resulting Nd<sup>3+</sup>→Yb<sup>3+</sup>→activator ET could extend the effective excitation bands for conventional Yb<sup>3+</sup>-sensitized UCNPs. Featuring lower water absorptions, these alternative excitation bands are expected to minimize the tissue overheating effect caused by NIR laser exposure (the blue line represents the absorption spectrum of water). Adapted with permission from reference [<a href="#B113-nanomaterials-04-00129" class="html-bibr">113</a>]. Copyright: American Chemical Society, 2013.</p>
Full article ">Figure 13
<p><span class="html-italic">In vitro</span> and <span class="html-italic">in vivo</span> heating effect induced by laser irradiation. (<b>a</b>,<b>b</b>) HEK (Human Embryonic Kidney) 293T cells after 5 min irradiation of 980 nm (<b>a</b>) and a 808 nm laser (<b>b</b>). Living cells and dead cells were stained with calcein AM (Acetomethoxy) and propidium iodide, respectively. (<b>c</b>,<b>d</b>) Infrared thermal image of a nude mouse during continuous (<b>c</b>) 980 nm laser irradiation for 50 s and (<b>d</b>) 808 nm laser irradiation for 300 s. Irradiation spots are denoted with the white arrows. Adapted with permission from reference [<a href="#B113-nanomaterials-04-00129" class="html-bibr">113</a>]. Copyright: American Chemical Society, 2013.</p>
Full article ">
4641 KiB  
Review
Numerical Modeling of Sub-Wavelength Anti-Reflective Structures for Solar Module Applications
by Katherine Han and Chih-Hung Chang
Nanomaterials 2014, 4(1), 87-128; https://doi.org/10.3390/nano4010087 - 29 Jan 2014
Cited by 89 | Viewed by 14859
Abstract
This paper reviews the current progress in mathematical modeling of anti-reflective subwavelength structures. Methods covered include effective medium theory (EMT), finite-difference time-domain (FDTD), transfer matrix method (TMM), the Fourier modal method (FMM)/rigorous coupled-wave analysis (RCWA) and the finite element method (FEM). Time-based solutions [...] Read more.
This paper reviews the current progress in mathematical modeling of anti-reflective subwavelength structures. Methods covered include effective medium theory (EMT), finite-difference time-domain (FDTD), transfer matrix method (TMM), the Fourier modal method (FMM)/rigorous coupled-wave analysis (RCWA) and the finite element method (FEM). Time-based solutions to Maxwell’s equations, such as FDTD, have the benefits of calculating reflectance for multiple wavelengths of light per simulation, but are computationally intensive. Space-discretized methods such as FDTD and FEM output field strength results over the whole geometry and are capable of modeling arbitrary shapes. Frequency-based solutions such as RCWA/FMM and FEM model one wavelength per simulation and are thus able to handle dispersion for regular geometries. Analytical approaches such as TMM are appropriate for very simple thin films. Initial disadvantages such as neglect of dispersion (FDTD), inaccuracy in TM polarization (RCWA), inability to model aperiodic gratings (RCWA), and inaccuracy with metallic materials (FDTD) have been overcome by most modern software. All rigorous numerical methods have accurately predicted the broadband reflection of ideal, graded-index anti-reflective subwavelength structures; ideal structures are tapered nanostructures with periods smaller than the wavelengths of light of interest and lengths that are at least a large portion of the wavelengths considered. Full article
(This article belongs to the Special Issue Nanomaterials in Energy Conversion and Storage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Example images of moth-eye structures found in nature. The scale bars are (<b>a</b>) 10 µm; (<b>b</b>–<b>g</b>) 500 nm. Example image of <span class="html-italic">Pieris napi</span> (<b>h</b>). <a href="#nanomaterials-04-00087-f001" class="html-fig">Figure 1</a>a–g reprinted with permission from reference [<a href="#B4-nanomaterials-04-00087" class="html-bibr">4</a>], Copyright 2006 Elsevier; <a href="#nanomaterials-04-00087-f001" class="html-fig">Figure 1</a>h reprinted with permission from reference [<a href="#B5-nanomaterials-04-00087" class="html-bibr">5</a>], Copyright 2004–2013 John Pickering.</p>
Full article ">Figure 2
<p>Orientation diagram of 2-D grating (3-D model, <b>a</b>) and 1-D grating (2-D model, <b>b</b>).</p>
Full article ">Figure 3
<p>Diagram of transverse electric (TE) and transverse magnetic (TM) incident light at a non-zero angle of incidence on an interface plane for angle of incidence (AOI) &lt; Brewster’s angle.</p>
Full article ">Figure 4
<p>Diffraction orders on a 1-D grating at normal incidence.</p>
Full article ">Figure 5
<p>Schematic of a graded index subwavelength structure (<b>a</b>); the effective index of refraction according to how the light would interact with the material (<b>b</b>); and a graph of an approximate effective index (<b>c</b>); The various regimes of optical behavior with grating sizes are shown in (<b>d</b>), indicating that the effective medium theories are applicable for gratings whose periods (p) are much smaller than the wavelength of light (λ).</p>
Full article ">Figure 6
<p>TEM images of moth eye nipple arrays (<b>a</b>,<b>b</b>) and (<b>c</b>) the effective index of refraction for three nipple types that exhibit graded index (gradient index of refraction, GRIN) behavior. Reprinted with permission from reference [<a href="#B6-nanomaterials-04-00087" class="html-bibr">6</a>], Copyright 2010 Elsevier; and reference [<a href="#B4-nanomaterials-04-00087" class="html-bibr">4</a>], Copyright 2006 The Royal Society.</p>
Full article ">Figure 7
<p>Simulation with effective medium theory (EMT) and measurement results of short circuit current from organic solar cells with no ARC, with a moth eye ARC, and with an idea GRIN structure ARC. Reprinted with permission from reference [<a href="#B20-nanomaterials-04-00087" class="html-bibr">20</a>], Copyright 2008 Elsevier.</p>
Full article ">Figure 8
<p>TE and TM polarization hitting a 1-D grating. The wavevector is indicated as <span class="html-italic">k</span>.</p>
Full article ">Figure 9
<p>Reflectivity of several GRIN structures over a range of heights, wavelengths, and angles of incidence. Reprinted with permission from reference [<a href="#B23-nanomaterials-04-00087" class="html-bibr">23</a>], Copyright 2007 Nature Publishing Group.</p>
Full article ">Figure 10
<p>Yee cell. Arrows indicate the direction of the <span class="html-italic">E</span> or <span class="html-italic">H</span> field that is calculated at each point.</p>
Full article ">Figure 11
<p>Time-based <span class="html-italic">E</span>-fields introduced as a plane wave (<b>a</b>) to simulate a range of frequencies (<b>b</b>).</p>
Full article ">Figure 12
<p>Reflectivity calculated from the finite-difference time-domain (FDTD) and transfer matrix method (TMM) methods compared to experimental data shows that the FDTD method is more accurate than the TMM method. Reprinted with permission from reference [<a href="#B36-nanomaterials-04-00087" class="html-bibr">36</a>], Copyright 2003 IEEE.</p>
Full article ">Figure 13
<p>Reflectivity results from graded-index films with an integral RI profile ( <span class="html-fig-inline" id="nanomaterials-04-00087-i021"> <img alt="Nanomaterials 04 00087 i021" src="/nanomaterials/nanomaterials-04-00087/article_deploy/html/images/nanomaterials-04-00087-i021.png"/></span>), square pyramids with linear and quantic RI profiles closely packed in a square lattice, and cones closely packed in a triangular lattice from EMT (lines) and FDTD (points) calculations. Reprinted with permission from reference [<a href="#B32-nanomaterials-04-00087" class="html-bibr">32</a>], Copyright 2011 Optical Society of America.</p>
Full article ">Figure 14
<p>Comparisons between reflectivity calculated by EMT, FDTD, and experimental results for GRIN structures. Reprinted with permission from reference [<a href="#B40-nanomaterials-04-00087" class="html-bibr">40</a>], Copyright 2007 Optical Society of America.</p>
Full article ">Figure 15
<p>SEM image of square packed silicon cones (<b>a</b>) and (<b>b</b>) comparison of FDTD (dotted) and experimental results (solid). Reprinted with permission from reference [<a href="#B32-nanomaterials-04-00087" class="html-bibr">32</a>], Copyright 2009 Optical Society of America.</p>
Full article ">Figure 16
<p>Angle of incidence FDTD simulations for nanorod arrays in TE and TM polarization [<a href="#B47-nanomaterials-04-00087" class="html-bibr">47</a>]. Reprinted with permission from reference [<a href="#B47-nanomaterials-04-00087" class="html-bibr">47</a>], Copyright 2011 AIP Publishing.</p>
Full article ">Figure 17
<p>Geometry of the transfer matrix method. Each layer left to right represents a separate, homogeneous material. <span class="html-italic">H</span> is the magnetic field and <span class="html-italic">E</span> is the electric field. The thickness, <span class="html-italic">t</span>, of each layer is drawn. The RI is labeled as <span class="html-italic">n</span>, reflected wave energy as ρ, transmitted energy as τ, and <span class="html-italic">k</span> is a designation of layer position (this diagram includes layers 1 through <span class="html-italic">k</span> + 1).</p>
Full article ">Figure 18
<p>SEM image of a seven layered thin film ARC produced and modeled by Kuo <span class="html-italic">et al</span>. [<a href="#B61-nanomaterials-04-00087" class="html-bibr">61</a>] using TMM. Reprinted with permission from reference [<a href="#B61-nanomaterials-04-00087" class="html-bibr">61</a>], Copyright 2008 Optical Society of America.</p>
Full article ">Figure 19
<p>Stair step approximation of a periodic geometry as drawn in the commercially available rigorous coupled-wave analysis (RCWA) software, GD-Calc. Reprinted with permission from reference [<a href="#B72-nanomaterials-04-00087" class="html-bibr">72</a>], Copyright 2006 AIP Publishing.</p>
Full article ">Figure 20
<p>Geometry for planar-grating diffraction. Region 1 (<b>left</b>) is the region of incident light, which will contain both incident and reflected waves. Region 2 (<b>middle</b>) is the transmission grating, which will contain transmitted and reflected waves. Region 2 (<b>right</b>) is the substrate, which will contain only forward diffracted (transmitted) waves. This example uses a simple transmission grating with a sinusoidal permittivity that is oriented at an arbitrary angle from normal.</p>
Full article ">Figure 21
<p>Several moth-eye-like structures with different refraction index profiles and their associated reflectance <span class="html-italic">vs</span>. period and height at a wavelength of 1000 nm. Reprinted with permission from reference [<a href="#B72-nanomaterials-04-00087" class="html-bibr">72</a>], Copyright 2008 AIP Publishing.</p>
Full article ">Figure 22
<p>Klopfenstein structures have much better AR properties at shorter heights than do pyramids. Klopfenstein structures are shown on the (<b>left</b>), and the RCWA-calculated reflectivity <span class="html-italic">vs</span>. normalized depth for Klopfenstein and pyramid structures is shown on the (<b>right</b>). Right side reprinted with permission from reference [<a href="#B97-nanomaterials-04-00087" class="html-bibr">97</a>], Copyright 1995 Optical Society of America.</p>
Full article ">Figure 23
<p>(<b>a</b>) Geometry for disordered GaN cones modeled as a 49-cone unit cell using RCWA. Measured and calculated reflectance for the disordered nanopillars shown in <a href="#nanomaterials-04-00087-f023" class="html-fig">Figure 23</a> at <span class="html-italic">s</span> (TE) and <span class="html-italic">p</span> (TM) polarizations show that RCWA is more accurate in TE mode than in TM mode for pillars of (<b>b</b>) 300 nm; (<b>c</b>) 550 nm; and (<b>d</b>) 720 nm. Reprinted with permission from reference [<a href="#B114-nanomaterials-04-00087" class="html-bibr">114</a>], Copyright 2008 Optical Society of America.</p>
Full article ">Figure 24
<p>RCWA simulations accurately predict experimental transmission results by modeling the SWS as a super-Gaussian profile with one dimensional height variations with a standard deviation of 15%. Reprinted with permission from reference [<a href="#B114-nanomaterials-04-00087" class="html-bibr">114</a>], Copyright 2010 Optical Society of America.</p>
Full article ">Figure 25
<p>Experimental and simulated results from flat (black) and etched nipple array (red) show good agreement for the RCWA method. Blue line is an experimental value from a commercial c-Si solar cell. Reprinted with permission from reference [<a href="#B81-nanomaterials-04-00087" class="html-bibr">81</a>], Copyright 2008 AIP Publishing.</p>
Full article ">Figure 26
<p>The thin film multilayer model (TMM) and RCWA show very similar results for moth eye structures with 210 nm bases and 800 nm heights. Reprinted with permission from reference [<a href="#B81-nanomaterials-04-00087" class="html-bibr">81</a>], Copyright 2008 AIP Publishing.</p>
Full article ">Figure 27
<p>Diagram of FEM simulation, including meshing scheme. Figure Reprinted with permission from reference [<a href="#B123-nanomaterials-04-00087" class="html-bibr">123</a>], Copyright 2007 Optical Society of America.</p>
Full article ">Figure 28
<p>Comparison of reflectance calculations from FEM and TMM. Reprinted with permission from reference [<a href="#B125-nanomaterials-04-00087" class="html-bibr">125</a>], Copyright 2010 Elsevier.</p>
Full article ">
1665 KiB  
Article
One-Pot Solvothermal Synthesis of Highly Emissive, Sodium-Codoped, LaF3 and BaLaF5 Core-Shell Upconverting Nanocrystals
by Joshua T. Stecher, Anne B. Rohlfing and Michael J. Therien
Nanomaterials 2014, 4(1), 69-86; https://doi.org/10.3390/nano4010069 - 8 Jan 2014
Cited by 16 | Viewed by 7882
Abstract
We report a one-pot solvothermal synthesis of sub-10 nm, dominant ultraviolet (UV) emissive upconverting nanocrystals (UCNCs), based on sodium-codoped LaF3 and BaLaF5 (0.5%Tm; 20%Yb) and their corresponding core@shell derivatives. Elemental analysis shows a Na-codopant in these crystal systems of ~20% the [...] Read more.
We report a one-pot solvothermal synthesis of sub-10 nm, dominant ultraviolet (UV) emissive upconverting nanocrystals (UCNCs), based on sodium-codoped LaF3 and BaLaF5 (0.5%Tm; 20%Yb) and their corresponding core@shell derivatives. Elemental analysis shows a Na-codopant in these crystal systems of ~20% the total cation content; X-ray diffraction (XRD) data indicate a shift in unit cell dimensions consistent with these small codopant ions. Similarly, X-ray photoelectron spectroscopic (XPS) analysis reveals primarily substitution of Na+ for La3+ ions (97% of total Na+ codopant) in the crystal system, and interstitial Na+ (3% of detected Na+) and La3+ (3% of detected La3+) present in (Na)LaF3 and only direct substitution of Na+ for Ba2+ in Ba(Na)LaF5. In each case, XPS analysis of La 3d lines show a decrease in binding energy (0.08–0.25 eV) indicating a reduction in local crystal field symmetry surrounding rare earth (R.E.3+) ions, permitting otherwise disallowed R.E. UC transitions to be enhanced. Studies that examine the impact of laser excitation power upon luminescence intensity were conducted over 2.5–100 W/cm2 range to elucidate UC mechanisms that populate dominant UV emitting states. Low power saturation of Tm3+ 3F3 and 3H4 states was observed and noted as a key initial condition for effective population of the 1D2 and 1I6 UV emitting states, via Tm-Tm cross-relaxation. 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>Transmission electron microscopy (TEM) images of 0.5%Tm, 20%Yb codoped Upconverting Nanocrystals (UCNCs) in host lattices: (<b>a</b>) LaF<sub>3</sub> (12.6 ± 2.5 nm); (<b>b</b>) (Na)LaF<sub>3</sub> (7.7 ± 1.9 nm); (<b>c</b>) (Na)LaF<sub>3</sub>@(Na)LaF<sub>3</sub> (22.7 ± 4.5 nm); (<b>d</b>) BaLaF<sub>5</sub> (6.0 ± 1.4 nm); (<b>e</b>) Ba(Na)LaF<sub>5</sub> (7.5 ± 2.3 nm); (<b>f</b>) Ba(Na)LaF<sub>5</sub>@Ba(Na)LaF<sub>3</sub> (9.6 ± 2.3 nm): (200 kx magnification; 50 nm scale bars).</p>
Full article ">Figure 2
<p>Comparative emission spectra of 0.5%Tm, 20%Yb codoped UCNCs in host lattices. (<b>a1</b>) LaF<sub>3</sub> (12.6 ± 2.5 nm) (intensity shown scaled ×1000) (<b>Green</b>); (<b>a2</b>) (Na)LaF<sub>3</sub> (7.7 ± 1.9 nm) (<b>Blue</b>); (<b>a3</b>) (Na)LaF<sub>3</sub>@(Na)LaF<sub>3</sub> (22.7 ± 4.5 nm) (<b>Violet</b>); (<b>b1</b>) BaLaF<sub>5</sub> (6.0 ± 1.4 nm) (intensity shown scaled ×10 (<b>Green</b>); (<b>b2</b>) Ba(Na)LaF<sub>5</sub> (7.5 ± 2.3 nm) (<b>Blue</b>); (<b>b3</b>) Ba(Na)LaF<sub>5</sub>@Ba(Na)LaF<sub>3</sub> (9.6 ± 2.3 nm) (<b>Violet</b>); experimental conditions: 1 mg/mL solutions in toluene; 980 nm CW laser excitation; 60 W/cm<sup>2</sup>. Tm<sup>3+</sup> transitions assigned.</p>
Full article ">Figure 3
<p>X-ray powder diffraction (XRD) data acquired for: (<b>a1</b>) LaF<sub>3</sub> (<b>Green</b>); (<b>a2</b>) (Na)LaF<sub>3</sub> (<b>Blue</b>); (<b>a3</b>) (Na)LaF<sub>3</sub>@(Na)LaF<sub>3</sub> (<b>Violet</b>); (<b>a4</b>) line spectrum of hexagonal LaF<sub>3</sub>, JCPDS: 72-1435 (<b>Black</b>); hexagonal NaYbF<sub>4</sub> inclusions denoted in (Na)LaF<sub>3</sub>@(Na)LaF<sub>3</sub>: (<b>b1</b>) BaLaF<sub>5</sub> (<b>Green</b>); (<b>b2</b>) Ba(Na)LaF<sub>5</sub> (<b>Blue</b>); (<b>b3</b>) Ba(Na)LaF<sub>5</sub>@Ba(Na)LaF<sub>3</sub> (<b>Violet</b>); (<b>b4</b>) displays JCPDS: 43-0394 (cubic BaCeF<sub>5</sub> comparative, <b>Black</b>); and (<b>a5</b>,<b>b5</b>) line spectrum of hexagonal NaLaF<sub>4</sub>, JCPDS: 75-1923 (<b>Red</b>).</p>
Full article ">Figure 4
<p>XPS survey scans of (<b>a1</b>) (Na)LaF<sub>3</sub> (<b>pale Blue</b>); (<b>a2</b>) LaF<sub>3</sub> (<b>pale Green</b>); (<b>a3</b>) Ba(Na)LaF<sub>5</sub> (<b>Blue</b>); and (<b>a4</b>) BaLaF<sub>5</sub> (<b>Green</b>) indicating the presence and absence of Na-codopant; normalized XPS La 3d spectral region (<b>b1</b>) (Na)LaF<sub>3</sub> (<b>pale Blue</b>); (<b>b2</b>) LaF<sub>3</sub> (<b>pale Green</b>); (<b>b3</b>) Ba(Na)LaF<sub>5</sub> (<b>Blue</b>); and (<b>b4</b>) BaLaF<sub>5</sub> (<b>Green)</b> indicating a shift in binding energy in La 3d<sub>5/2</sub> core orbital upon Na<sup>+</sup> inclusion.</p>
Full article ">Figure 5
<p>Emission intensity <span class="html-italic">vs</span>. 980 nm laser power density (W/cm<sup>2</sup>), ln-ln IvP plot of 0.5%Tm, 20%Yb codoped (Na)LaF<sub>3</sub> UCNC in (<b>a</b>) UV/blue and (<b>b</b>) NIR regime, noting low power saturation of Tm<sup>3+</sup> <sup>3</sup>F<sub>3</sub> and <sup>3</sup>H<sub>4</sub> states; (<b>c</b>) Emission spectrum at 20 W/cm<sup>2</sup> and (<b>d</b>) ETU transition diagram for Yb<sup>3+</sup>-Tm<sup>3+</sup>. Experimental conditions: 1 mg/mL solutions in toluene at 23 °C; 980 nm CW laser excitation varying from 2.5 to 100 W/cm<sup>2</sup>.</p>
Full article ">Figure 6
<p>Emission intensity <span class="html-italic">vs</span>. 980 nm laser power density (W/cm<sup>2</sup>), expressed in ln-ln IvP plots of 0.5% Tm, 20% Yb codoped (Na)LaF<sub>3</sub> core@shell in the (<b>a</b>) UV/blue and (<b>b</b>) NIR regimes, noting low power saturation of Tm<sup>3+</sup> <sup>3</sup>H<sub>4</sub> state and onset of saturation (diamond markers); (<b>c</b>) Emission spectrum recorded at 20 W/cm<sup>2</sup> and (<b>d</b>) ETU transition diagram of Yb<sup>3+</sup>-Tm<sup>3+</sup>. Experimental conditions: 1 mg/mL solutions in toluene at 23 °C; 980 nm CW laser excitation varying from 2.5 to 100 W/cm<sup>2</sup>.</p>
Full article ">
2535 KiB  
Article
Enhanced Upconversion Luminescence in Yb3+/Tm3+-Codoped Fluoride Active Core/Active Shell/Inert Shell Nanoparticles through Directed Energy Migration
by Hailong Qiu, Chunhui Yang, Wei Shao, Jossana Damasco, Xianliang Wang, Hans Ågren, Paras N. Prasad and Guanying Chen
Nanomaterials 2014, 4(1), 55-68; https://doi.org/10.3390/nano4010055 - 3 Jan 2014
Cited by 76 | Viewed by 12380
Abstract
The luminescence efficiency of lanthanide-doped upconversion nanoparticles is of particular importance for their embodiment in biophotonic and photonic applications. Here, we show that the upconversion luminescence of typically used NaYF4:Yb3+30%/Tm3+0.5% nanoparticles can be enhanced by ~240 times [...] Read more.
The luminescence efficiency of lanthanide-doped upconversion nanoparticles is of particular importance for their embodiment in biophotonic and photonic applications. Here, we show that the upconversion luminescence of typically used NaYF4:Yb3+30%/Tm3+0.5% nanoparticles can be enhanced by ~240 times through a hierarchical active core/active shell/inert shell (NaYF4:Yb3+30%/Tm3+0.5%)/NaYbF4/NaYF4 design, which involves the use of directed energy migration in the second active shell layer. The resulting active core/active shell/inert shell nanoparticles are determined to be about 11 times brighter than that of well-investigated (NaYF4:Yb3+30%/Tm3+0.5%)/NaYF4 active core/inert shell nanoparticles when excited at ~980 nm. The strategy for enhanced upconversion in Yb3+/Tm3+-codoped NaYF4 nanoparticles through directed energy migration might have implications for other types of lanthanide-doped upconversion nanoparticles. 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>Transmission electron images (TEM) of: (<b>a</b>) The core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup> nanoparticles; (<b>b</b>) The active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub> nanoparticles; (<b>c</b>) The active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub> nanoparticles; and (<b>d</b>) The active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles. Histogram of size distribution of (<b>e</b>) The core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup> nanoparticles; (<b>f</b>) The active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub> nanoparticles; (<b>g</b>) The active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub> nanoparticles; and (<b>h</b>) The active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles. The size was evaluated according to TEM images of corresponding nanoparticles dispersed in hexane at a concentration of 0.1 wt.%.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of the core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup> nanoparticles, the active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub> nanoparticles, the active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub> nanoparticles, and the active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles, contrasted with the standard cubic NaYF<sub>4</sub> structure of JCPDS 06-0342.</p>
Full article ">Figure 3
<p>(<b>a</b>) Compared upconversion luminescence of colloidal nanoparticles (hexane dispersion, 1 wt.%) of the core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>, the active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub>, and the active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>. The peaks at ~690 and ~720 nm marked by asterisk correspond to the second order of the peaks at 345 and 360 nm; (<b>b</b>) Compared upconversion luminescence of colloidal nanoparticles (hexane dispersion, 1 wt.%) of the active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> and the active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>. The same batch of the core nanoparticles is utilized to grow the active core/inert shell, the active core/active shell, and the active core/active shell/inert shell nanoparticles. The excitation is performed with a diode laser at ~980 nm of about 50 W/cm<sup>2</sup>. All the spectra have been calibrated by the spectral sensitivity of the utilized spectrophotometer system.</p>
Full article ">Figure 4
<p>(<b>a</b>) A log-log plot of the dependence of various luminescence intensities from the active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles on the excitation density; (<b>b</b>) Energy level diagrams of Yb<sup>3+</sup> and Tm<sup>3+</sup> ions as well as the involved mechanisms for upconversion luminescence from different energy states of Tm<sup>3+</sup> ions.</p>
Full article ">Figure 5
<p>Decays of upconversion luminescence at 802 nm from the core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>, the active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub>, the active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>, and the active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles.</p>
Full article ">Figure 6
<p>Schematic illustration of the luminescence quenching mechanism in: (<b>a</b>) The core NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup> nanoparticles, as well as the luminescence enhancement mechanism in (<b>b</b>) The active core/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYF<sub>4</sub> nanoparticles; (<b>c</b>) The active core/active shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub> nanoparticles; and (<b>d</b>) The active core/active shell/inert shell (NaYF<sub>4</sub>: 30%Yb<sup>3+</sup>, 0.5%Tm<sup>3+</sup>)/NaYbF<sub>4</sub>/NaYF<sub>4</sub> nanoparticles.</p>
Full article ">
963 KiB  
Article
Effect of Low-Frequency AC Magnetic Susceptibility and Magnetic Properties of CoFeB/MgO/CoFeB Magnetic Tunnel Junctions
by Yuan-Tsung Chen, Sung-Hao Lin and Tzer-Shin Sheu
Nanomaterials 2014, 4(1), 46-54; https://doi.org/10.3390/nano4010046 - 2 Jan 2014
Cited by 6 | Viewed by 6687
Abstract
In this investigation, the low-frequency alternate-current (AC) magnetic susceptibility (χac) and hysteresis loop of various MgO thickness in CoFeB/MgO/CoFeB magnetic tunneling junction (MTJ) determined coercivity (Hc) and magnetization (Ms) and correlated that with χac [...] Read more.
In this investigation, the low-frequency alternate-current (AC) magnetic susceptibility (χac) and hysteresis loop of various MgO thickness in CoFeB/MgO/CoFeB magnetic tunneling junction (MTJ) determined coercivity (Hc) and magnetization (Ms) and correlated that with χac maxima. The multilayer films were sputtered onto glass substrates and the thickness of intermediate barrier MgO layer was varied from 6 to 15 Å. An experiment was also performed to examine the variation of the highest χac and maximum phase angle (θmax) at the optimal resonance frequency (fres), at which the spin sensitivity is maximal. The results reveal that χac falls as the frequency increases due to the relationship between magnetization and thickness of the barrier layer. The maximum χac is at 10 Hz that is related to the maximal spin sensitivity and that this corresponds to a MgO layer of 11 Å. This result also suggests that the spin sensitivity is related to both highest χac and maximum phase angle. The corresponding maximum of χac is related to high exchange coupling. High coercivity and saturation magnetization contribute to high exchange-coupling χac strength. Full article
(This article belongs to the Special Issue Magnetic Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Measured low-frequency alternate-current magnetic susceptibility (χ<sub>ac</sub>) of CoFeB/MgO/CoFeB as a function of thickness of MgO barrier layer.</p>
Full article ">Figure 2
<p>Maximum χ<sub>ac</sub> as a function of thickness of MgO barrier layer.</p>
Full article ">Figure 3
<p>Variation of maximum χ<sub>ac</sub> with maximum phase angle (θ).</p>
Full article ">Figure 4
<p>The essential magnetic properties of magnetic tunneling junction are (<b>a</b>) hysteresis loop of MTJ, (<b>b</b>) coercivity value, and (<b>c</b>) saturation magnetization.</p>
Full article ">
1047 KiB  
Review
Multiple Exciton Generation in Colloidal Nanocrystals
by Charles Smith and David Binks
Nanomaterials 2014, 4(1), 19-45; https://doi.org/10.3390/nano4010019 - 24 Dec 2013
Cited by 81 | Viewed by 16186
Abstract
In a conventional solar cell, the energy of an absorbed photon in excess of the band gap is rapidly lost as heat, and this is one of the main reasons that the theoretical efficiency is limited to ~33%. However, an alternative process, multiple [...] Read more.
In a conventional solar cell, the energy of an absorbed photon in excess of the band gap is rapidly lost as heat, and this is one of the main reasons that the theoretical efficiency is limited to ~33%. However, an alternative process, multiple exciton generation (MEG), can occur in colloidal quantum dots. Here, some or all of the excess energy is instead used to promote one or more additional electrons to the conduction band, potentially increasing the photocurrent of a solar cell and thereby its output efficiency. This review will describe the development of this field over the decade since the first experimental demonstration of multiple exciton generation, including the controversies over experimental artefacts, comparison with similar effects in bulk materials, and the underlying mechanisms. We will also describe the current state-of-the-art and outline promising directions for further development. Full article
(This article belongs to the Special Issue Nanomaterials in Energy Conversion and Storage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photogeneration of charges in a (<b>A</b>) bulk semiconductor and (<b>B</b>) a nanocrystal quantum dot. In (<b>B</b>) quantum confinement effects increase the separation between energy levels, inhibiting phonon cooling, enhancing the multiple exciton generation (MEG) process and reducing waste heat.</p>
Full article ">Figure 2
<p>Schematic of a typical ultrafast transient absorption (UTA) experiment for measuring MEG. A Ti:Sapphire regenerative amplifier seeded by a Ti:Sapphire oscillator produces a pulsed beam at 800 nm, with 100 fs pulse width, ~1 mJ pulse energy and 1 kHz repetition rate (red line). Using a beam splitter 95% of this beam is passed to an optical parametric amplifier (OPA) with harmonic generating (HG) crystals to produce a pump beam that is tunable from the infra-red to the ultra-violet (blue line). The pump beam is passed through a mechanical chopper to improve the signal-to-noise ratio and then focused onto the sample. The remaining 5% of the beam from the amplifier is directed through a delay stage to vary the arrival time difference between pump and probe pulses at the sample. This beam is then passed through a sapphire plate to produce a white light continuum, which is split to form the probe and reference beams (yellow line). After the sample, the probe and reference beams are balanced using a neutral density (ND) filter (in the absence of the pump), passed through a spectrometer, then onto separate photodetectors. Small differences in probe and reference are detected using a lock-in amplifier synchronised to the chopper.</p>
Full article ">Figure 3
<p>(<b>a</b>) A typical signal obtained from a UTA experiment showing the fractional transmittance change, Δ<span class="html-italic">T</span>/<span class="html-italic">T</span>, as a function of time delay between the pump and probe beams. <span class="html-italic">R</span> is the ratio between the peak amplitude, <span class="html-italic">A</span>, and the plateau amplitude, <span class="html-italic">B</span>, where enough time has passed that all bi-excitons have decayed, but before any significant decay of single-excitons; (<b>b</b>) <span class="html-italic">R</span> as a function of fractional transmittance change, Δ<span class="html-italic">T</span>/<span class="html-italic">T</span>, for InP QD excited at 2.6 times <span class="html-italic">E</span><sub>g</sub>. The line is a fit to Equation (2), data from Reference [<a href="#B24-nanomaterials-04-00019" class="html-bibr">24</a>]; (<b>c</b>) Additional excitons produced, (QY − 1), as a function of <span class="html-italic">hv</span>/<span class="html-italic">E</span><sub>g</sub> for InAs NQD of different <span class="html-italic">E</span><sub>g</sub>. <span class="html-italic">hv</span><sub>th</sub> corresponds to the point where the rise in additional excitons begins, data from Reference [<a href="#B25-nanomaterials-04-00019" class="html-bibr">25</a>]. Reproduced from Reference [<a href="#B26-nanomaterials-04-00019" class="html-bibr">26</a>] by permission of the PCCP Owner Societies.</p>
Full article ">Figure 4
<p>Trion formation in a QD provides a charge capable of receiving and dissipating the energy liberated in the recombination of an exciton, resulting in an enhanced single-exciton recombination rate.</p>
Full article ">Figure 5
<p>Models of biexciton formation from a single absorbed photon: (<b>a</b>) a photo-generated hot exciton relaxes by impact ionization, exciting another electron across the band gap; thereby creating a biexciton (<b>b</b>) the direct photogeneration of a biexciton through a virtual exciton or biexciton states [<a href="#B62-nanomaterials-04-00019" class="html-bibr">62</a>]; and (<b>c</b>) the photo-generation, by a single photon, of a quantum superposition of all possible excited states, including multi-excitons.</p>
Full article ">Figure 6
<p>(<b>a</b>) Energy division in a QD sample exhibiting ideal MEG modeled using Equation (12) if illuminated by the AM1.5 solar spectrum. As the bandgap increases a larger fraction of solar photons do not have enough energy to be absorbed, indicated by the blue section, and for lower bandgap the threshold for MEG is lower; (<b>b</b>) Comparison of current density as a function of operating voltage for QD solar cells of optimum <span class="html-italic">E</span><sub>g</sub> with and without ideal MEG; (<b>c</b>) Comparison of solar cell efficiency <span class="html-italic">vs.</span> <span class="html-italic">E</span><sub>g</sub> simulated using Equation (13), for PbSe QD with experimentally derived η and <span class="html-italic">hv</span><sub>th</sub>, to the ideal case from (<b>a</b>) and the case with no MEG.</p>
Full article ">Figure 7
<p>(<b>a</b>) Layer structure of a Grätzel-type photovoltaic device utilising PbSe QDs as the absorbing species, with charge carrier transfer illustrated [<a href="#B77-nanomaterials-04-00019" class="html-bibr">77</a>]; (<b>b</b>) Layer structure of a depleted heterojunction design photovoltaic device utilising PbSe QDs as the absorbing species [<a href="#B78-nanomaterials-04-00019" class="html-bibr">78</a>].</p>
Full article ">Figure 8
<p>Illustration of the processes which compete with MEG, yellow arrows represent waste energy. (<b>a</b>) Phonon emission; (<b>b</b>) Electron transfer to a surface state; (<b>c</b>) Auger relaxation, where the electron’s excess energy is transferred to the hole which exhibits faster phonon cooling due to the denser band structure of the valence band; (<b>d</b>) Transfer of energy to surface ligands through vibrational coupling.</p>
Full article ">
1899 KiB  
Article
Composite Electrolyte Membranes from Partially Fluorinated Polymer and Hyperbranched, Sulfonated Polysulfone
by Surya Subianto, Namita Roy Choudhury and Naba Dutta
Nanomaterials 2014, 4(1), 1-18; https://doi.org/10.3390/nano4010001 - 23 Dec 2013
Cited by 30 | Viewed by 9898
Abstract
Macromolecular modification of poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF) was done with various proportions of sulfonic acid terminated, hyperbranched polysulfone (HPSU) with a view to prepare ion conducting membranes. The PVDF-co-HFP was first chemically modified by dehydrofluorination and chlorosulfonation in order to make the membrane more [...] Read more.
Macromolecular modification of poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF) was done with various proportions of sulfonic acid terminated, hyperbranched polysulfone (HPSU) with a view to prepare ion conducting membranes. The PVDF-co-HFP was first chemically modified by dehydrofluorination and chlorosulfonation in order to make the membrane more hydrophilic as well as to introduce unsaturation, which would allow crosslinking of the PVDF-co-HFP matrix to improve the stability of the membrane. The modified samples were characterized for ion exchange capacity, morphology, and performance. The HPSU modified S-PVDF membrane shows good stability and ionic conductivity of 5.1 mS cm1 at 80 °C and 100% RH for blends containing 20% HPSU, which is higher than the literature values for equivalent blend membranes using Nafion. SEM analysis of the blend membranes containing 15% or more HPSU shows the presence of spherical domains with a size range of 300–800 nm within the membranes, which are believed to be the HPSU-rich area. Full article
(This article belongs to the Special Issue Nanomaterials in Energy Conversion and Storage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>MALDI spectrum of the synthesized HPSU.</p>
Full article ">Figure 2
<p>PA-FTIR spectra of the HPSU and its precursor compounds.</p>
Full article ">Figure 3
<p>TGA comparison of HPSU and its precursor compound.</p>
Full article ">Figure 4
<p>Photoacoustic FTIR spectra comparison between PVDF-co-HFP, dehydrofluorinated PVDF-co-HFP, and S-PVDF.</p>
Full article ">Figure 5
<p>High resolution XPS C1s spectrum of PVDF-co-HFP, dehydrofluorinated PVDF-co-HFP, and S-PVDF.</p>
Full article ">Figure 6
<p>PA-FTIR Spectra of S-PVDF and the S-PVDF/HPSU composite.</p>
Full article ">Figure 7
<p>Cross section SEM images of the S-PVDF/HPSU composite membranes with 10%, 15% (volume %) HPSU (magnifications: left images 20 μm; right images: magnified view at 2 μm). Sample cross section was obtained by fracturing the membranes in liquid N<sub>2</sub>.</p>
Full article ">Figure 8
<p>TGA spectra of the blend membrane (20% HPSU) and untreated PVDF-co-HFP.</p>
Full article ">Figure 9
<p>DSC thermograms of the blend membranes (10%, 15%, 20% HPSU) with S-PVDF.</p>
Full article ">Figure 10
<p>Storage moduli change of different blends at a constant humidity of 50%.</p>
Full article ">Figure 11
<p>Ionic conductivity at 80°C and 100% humidity and water uptake of samples containing 10%–20% HPSU.</p>
Full article ">Figure 12
<p>Synthesis of the sulfonic acid terminated, hyperbranched polysulfone.</p>
Full article ">Figure 13
<p>Sulfonation of PVDF-co-HFP.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop