[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 8, May
Previous Issue
Volume 8, March
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
polymers-logo

Journal Browser

Journal Browser

Polymers, Volume 8, Issue 4 (April 2016) – 65 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
3837 KiB  
Article
Smart Homopolymer Microgels: Influence of the Monomer Structure on the Particle Properties
by Bastian Wedel, Yvonne Hertle, Oliver Wrede, Johannes Bookhold and Thomas Hellweg
Polymers 2016, 8(4), 162; https://doi.org/10.3390/polym8040162 - 23 Apr 2016
Cited by 53 | Viewed by 11316
Abstract
In this work, we compare the properties of smart homopolymer microgels based on N-n-propylacrylamide (NNPAM), N-isopropylacrylamide (NIPAM) and N-isopropylmethacrylamide (NIPMAM) synthesized under identical conditions. The particles are studied with respect to size, morphology, and swelling behavior using scanning electron and [...] Read more.
In this work, we compare the properties of smart homopolymer microgels based on N-n-propylacrylamide (NNPAM), N-isopropylacrylamide (NIPAM) and N-isopropylmethacrylamide (NIPMAM) synthesized under identical conditions. The particles are studied with respect to size, morphology, and swelling behavior using scanning electron and scanning force microscopy. In addition, light scattering techniques and fluorescent probes are employed to follow the swelling/de-swelling of the particles. Significant differences are found and discussed. Poly(N-n-propylacrylamide) (PNNPAM) microgels stand out due to their very sharp volume phase transition, whereas Poly(N-isopropylmethacrylamide) (PNIPMAM) particles are found to exhibit a more homogeneous network structure compared to the other two systems. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images of the homopolymer microgels of (<b>a</b>) NNPAM, (<b>b</b>) NIPAM and (<b>c</b>) NIPMAM synthesized without surfactant and the corresponding size distributions (<b>d</b>–<b>f</b>). The black lines represent a Gaussian distribution function.</p>
Full article ">Figure 2
<p>AFM images of the homopolymers of (<b>a</b>) NNPAM, (<b>b</b>) NIPAM and (<b>c</b>) NIPMAM synthesized without surfactant and the corresponding size distributions (<b>d</b>–<b>f</b>). The black lines represent a Gaussian distribution function.</p>
Full article ">Figure 3
<p>From left to right: height profiles of (<b>a</b>) NNPAM, (<b>b</b>) NIPAM and (<b>c</b>) NIPMAM homopolymer microgels obtained from AFM measurements in tapping mode. Each line represents one microgel particle of the same sample. The samples are characterized in the dried state.</p>
Full article ">Figure 4
<p>Phase images of (<b>a</b>) PNNPAM, (<b>b</b>) PNIPAM and (<b>c</b>) PNIPMAM homopolymer microgels obtained from AFM measurements.</p>
Full article ">Figure 5
<p>Plot of the absolute scattering intensities vs. the scattering vector <span class="html-italic">q</span> of PNNPAM, PNIPAM and PNIPMAM at 15 <math display="inline"> <semantics> <msup> <mrow/> <mo>°</mo> </msup> </semantics> </math>C and the corresponding fit with the fuzzy sphere model (<b>left</b>) [<a href="#B17-polymers-08-00162" class="html-bibr">17</a>]. For a better representation of the experimental data, the intensity values were shifted along the <span class="html-italic">y</span>-axis. Comparison of the homogeneous [<a href="#B39-polymers-08-00162" class="html-bibr">39</a>] and the fuzzy sphere form factor model on the example of PNIPMAM (<b>right</b>).</p>
Full article ">Figure 6
<p>Structural formulas of the various acrylamide radicals during polymerisation (left to right: NNPAM, NIPAM and NIPMAM). In NIPMAM, the radical is stabilized by the methyl group and its positive inductive effect, whereby the polymerization rate is reduced.</p>
Full article ">Figure 7
<p>(<b>a</b>) PCS relaxation rate distribution of the collapsed homopolymer microgels; (<b>b</b>) Plot of the mean relaxation rate <math display="inline"> <semantics> <mover> <mo>Γ</mo> <mo>¯</mo> </mover> </semantics> </math> <span class="html-italic">vs.</span> <math display="inline"> <semantics> <msup> <mi>q</mi> <mn>2</mn> </msup> </semantics> </math> and the corresponding linear regression to determine the translational diffusion coefficient <math display="inline"> <semantics> <msup> <mi>D</mi> <mi>T</mi> </msup> </semantics> </math>.</p>
Full article ">Figure 8
<p>The swelling curves show the change of the microgel hydrodynamic radius (<b>a</b>) and the swelling ratio <span class="html-italic">α</span> (<b>b</b>) as a function of temperature for all three homopolymers. The graphs (<b>c</b>) and (<b>d</b>) represent the numerically calculated first derivative of the swelling curves. The point of inflection is the volume phase transition temperature (VPTT) of the system.</p>
Full article ">Figure 9
<p>Normalized light attenuation <math display="inline"> <semantics> <msub> <mi>δ</mi> <mn>700</mn> </msub> </semantics> </math> at <span class="html-italic">λ</span> = 700 nm for PNNPAM (<b>black curve</b>), PNIPAM (<b>red curve</b>) and PNIPMAM microgels (<b>blue curve</b>) as a function of temperature measured by UV/Vis spectroscopy.</p>
Full article ">Figure 10
<p>Analysis of the phase transition of PNNPAM, PNIPAM and PNIPMAM microgels using UV/Vis spectroscopy. (<b>a</b>) first derivative of the light attenuation coefficients with respect to temperature as a function of temperature; (<b>b</b>) detailed image of the temperature range between 25 <math display="inline"> <semantics> <msup> <mrow/> <mo>°</mo> </msup> </semantics> </math>C and 55 <math display="inline"> <semantics> <msup> <mrow/> <mo>°</mo> </msup> </semantics> </math>C relevant for PNIPAM and PNIPMAM particles.</p>
Full article ">Figure 11
<p>Normalized light attenuation coefficients for PNNPAM (<b>a</b>) and PNIPAM (<b>b</b>) microgels solutions with different concentrations as a function of temperature (<span class="html-italic">λ</span> = 700 nm). In graphs (<b>c</b>) and (<b>d</b>), a summary of the VPTTs and the <math display="inline"> <semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics> </math>-values for both microgels as a function of sample concentration is given. The solid lines correspond to the average value. It has to be mentioned that for PNIPAM the first <math display="inline"> <semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics> </math>-value (0.0015 wt %) has been neglected.</p>
Full article ">Figure 12
<p>(<b>a</b>) Plot of the intensity ratio <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>I</mi> <mn>3</mn> </msub> </mrow> </semantics> </math> as a function of temperature for PNNPAM, PNIPAM and NIPMAM solutions mixed with pyrene as well as pure pyrene as reference; (<b>b</b>) To quantify the VPTT, the first derivative of <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>I</mi> <mn>3</mn> </msub> </mrow> </semantics> </math> with respect to temperature as a function of temperature is presented. The concentration of all microgel solutions was 0.1 wt %.</p>
Full article ">Figure 13
<p>Plot of the intensity ratio <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>I</mi> <mn>3</mn> </msub> </mrow> </semantics> </math> as a function of temperature for PNNPAM microgel solutions with different concentrations (<b>left</b>). Additionally, the concentration dependent change in <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>I</mi> <mn>3</mn> </msub> </mrow> </semantics> </math> for the swollen and collapsed microgel is shown (<b>right</b>).</p>
Full article ">
5270 KiB  
Article
Stereocomplex-Reinforced PEGylated Polylactide Micelle for Optimized Drug Delivery
by Chunsheng Feng, Meihua Piao and Di Li
Polymers 2016, 8(4), 165; https://doi.org/10.3390/polym8040165 - 22 Apr 2016
Cited by 18 | Viewed by 6879
Abstract
The instability of PEGylated polylactide micelles is a challenge for drug delivery. Stereocomplex interaction between racemic polylactide chains with different configurations provides an effective strategy to enhance the stability of micelles as the nanocarriers of drugs. In this work, a stereocomplex micelle (SCM) [...] Read more.
The instability of PEGylated polylactide micelles is a challenge for drug delivery. Stereocomplex interaction between racemic polylactide chains with different configurations provides an effective strategy to enhance the stability of micelles as the nanocarriers of drugs. In this work, a stereocomplex micelle (SCM) self-assembled from the amphiphilic triblock copolymers comprising poly(ethylene glycol) (PEG), and dextrorotatory and levorotatory polylactides (PDLA and PLLA) was applied for efficient drug delivery. The spherical SCM showed the smallest scale and the lowest critical micelle concentration (CMC) than the micelles with single components attributed to the stereocomplex interaction between PDLA and PLLA. 10-Hydroxycamptothecin (HCPT) as a model antitumor drug was loaded into micelles. Compared with the loading micelles from individual PDLA and PLLA, the HCPT-loaded SCM exhibited the highest drug loading efficiency (DLE) and the slowest drug release in phosphate-buffered saline (PBS) at pH 7.4, indicating its enhanced stability in circulation. More fascinatingly, the laden SCM was demonstrated to have the highest cellular uptake of HCPT and suppress malignant cells most effectively in comparison to the HCPT-loaded micelles from single copolymer. In summary, the stereocomplex-enhanced PLA–PEG–PLA micelle may be promising for optimized drug delivery in the clinic. Full article
(This article belongs to the Special Issue Functional Polymers for Medical Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>1</sup>H NMR spectra (in CDCl<sub>3</sub>) (<b>A</b>) and GPC chromatograms (<b>B</b>) of PDLA–PEG–PDLA and PLLA–PEG–PLLA copolymers.</p>
Full article ">Figure 2
<p>Typical TEM micrographs and <span class="html-italic">D</span><sub>h</sub>s of PDM, PLM, and SCM.</p>
Full article ">Figure 3
<p>Crystallization peak variations of PDM, PLM, and SCM at ranges of 830–980 cm<sup>−1</sup> (<b>A</b>) and 1530–2070 cm<sup>−1</sup> (<b>B</b>).</p>
Full article ">Figure 4
<p>DSC thermal graphs of PDM, PLM, and SCM during heating (<b>A</b>) and cooling processes (<b>B</b>).</p>
Full article ">Figure 5
<p>WAXD profiles of PDM, PLM, and SCM.</p>
Full article ">Figure 6
<p>Release behaviors of PDM/HCPT, PLM/HCPT, and SCM/HCPT in PBS at pH 7.4, 37 °C. Each set of data is presented as mean ± standard deviation (SD; <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>Cellular uptake of HCPT toward MCF-7 cells after incubation with PDM/HCPT, PLM/HCPT, SCM/HCPT, or free HCPT for various durations. Data are shown as mean ± SD (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 8
<p>Relative viabilities of MCF-7 cells after incubation with blank micelles for 72 h (<b>A</b>); or HCPT-loaded micelles or free HCPT for 48 h (<b>B</b>) or 72 h (<b>C</b>). Data are shown as mean ± SD (<span class="html-italic">n</span> = 4).</p>
Full article ">Scheme 1
<p>Schematic illustration for preparation, endocytosis, and intracellular drug release of SCM/HCPT.</p>
Full article ">
5061 KiB  
Article
Modelling and Validation of Synthesis of Poly Lactic Acid Using an Alternative Energy Source through a Continuous Reactive Extrusion Process
by Satya P. Dubey, Hrushikesh A. Abhyankar, Veronica Marchante, James L. Brighton, Kim Blackburn, Clive Temple, Björn Bergmann, Giang Trinh and Chantal David
Polymers 2016, 8(4), 164; https://doi.org/10.3390/polym8040164 - 22 Apr 2016
Cited by 23 | Viewed by 8240
Abstract
PLA is one of the most promising bio-compostable and bio-degradable thermoplastic polymers made from renewable sources. PLA is generally produced by ring opening polymerization (ROP) of lactide using the metallic/bimetallic catalyst (Sn, Zn, and Al) or other organic catalysts in a suitable solvent. [...] Read more.
PLA is one of the most promising bio-compostable and bio-degradable thermoplastic polymers made from renewable sources. PLA is generally produced by ring opening polymerization (ROP) of lactide using the metallic/bimetallic catalyst (Sn, Zn, and Al) or other organic catalysts in a suitable solvent. In this work, reactive extrusion experiments using stannous octoate Sn(Oct)2 and tri-phenyl phosphine (PPh)3 were considered to perform ROP of lactide. Ultrasound energy source was used for activating and/or boosting the polymerization as an alternative energy (AE) source. Ludovic® software, designed for simulation of the extrusion process, had to be modified in order to simulate the reactive extrusion of lactide and for the application of an AE source in an extruder. A mathematical model for the ROP of lactide reaction was developed to estimate the kinetics of the polymerization process. The isothermal curves generated through this model were then used by Ludovic software to simulate the “reactive” extrusion process of ROP of lactide. Results from the experiments and simulations were compared to validate the simulation methodology. It was observed that the application of an AE source boosts the polymerization of lactide monomers. However, it was also observed that the predicted residence time was shorter than the experimental one. There is potentially a case for reducing the residence time distribution (RTD) in Ludovic® due to the ‘liquid’ monomer flow in the extruder. Although this change in parameters resulted in validation of the simulation, it was concluded that further research is needed to validate this assumption. Full article
(This article belongs to the Special Issue Computational Modeling and Simulation in Polymer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Methodology to explain reaction parameters. <span class="html-italic">M</span><sub>0</sub>: Initial concentration of monomer, <span class="html-italic">C</span><sub>0</sub>: Initial concentration of catalyst Sn(Oct)<sub>2</sub>, <span class="html-italic">D</span><sub>0</sub>: Initial concentration of OH group source, <span class="html-italic">A</span><sub>0</sub>: Initial concentration of phosphine (PPh)<sub>3</sub>, <math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <mo stretchy="false">(</mo> <msub> <mi>M</mi> <mi>n</mi> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> </mrow> </semantics> </math>) number average molecular weight, <span class="html-italic">X</span>: conversion, RT: residential time, <span class="html-italic">K</span>i: rate constants.</p>
Full article ">Figure 2
<p>Input and output for continuous extrusion simulation.</p>
Full article ">Figure 3
<p>Screw profile with AE source device.</p>
Full article ">Figure 4
<p>Temperature variation due to conduction (<b>purple</b>), due to mechanical effect (<b>red</b>) and micro-wave (AE) (100 W) (<b>green</b>) effect and result on final temperature (<b>grey</b>).</p>
Full article ">Figure 5
<p>Isothermal curves for conversion (<span class="html-italic">X</span>) <span class="html-italic">vs. t</span>.</p>
Full article ">Figure 6
<p>Isothermal curves for average molecular weight (<math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>M</mi> <mi>n</mi> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> </mrow> </semantics> </math>) <span class="html-italic">vs.</span> <span class="html-italic">t</span>.</p>
Full article ">Figure 7
<p><math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>M</mi> <mi mathvariant="normal">n</mi> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> </mrow> </semantics> </math> (<b>purple line</b>) <span class="html-italic">vs.</span> <span class="html-italic">T</span> &amp; <span class="html-italic">X</span> (<b>red line</b>) <span class="html-italic">vs. T</span> obtained with Ludovic<sup>®</sup> for <span class="html-italic">T</span> (50–220) °C, AE = 250 W, 600 rpm.</p>
Full article ">Figure 8
<p>RTD Comparison for Ludovic simulated (<b>red</b> dots) and experimental (<b>blue</b> dots) results for throughput and barrel temperature.</p>
Full article ">Figure 9
<p>Trial-2 RTD Comparison considering reaction at the middle stage of the extruder.</p>
Full article ">Figure 10
<p>RTD Comparison for Ludovic simulated (<b>red</b> dots) and experimental (<b>blue</b> dots) results for throughput and barrel temperature.</p>
Full article ">
2588 KiB  
Article
Biodegradable Polyphosphazene Based Peptide-Polymer Hybrids
by Anne Linhardt, Michael König, Wolfgang Schöfberger, Oliver Brüggemann, Alexander K. Andrianov and Ian Teasdale
Polymers 2016, 8(4), 161; https://doi.org/10.3390/polym8040161 - 22 Apr 2016
Cited by 30 | Viewed by 10342
Abstract
A novel series of peptide based hybrid polymers designed to undergo enzymatic degradation is presented, via macrosubstitution of a polyphosphazene backbone with the tetrapeptide Gly-Phe-Leu-Gly. Further co-substitution of the hybrid polymers with hydrophilic polyalkylene oxide Jeffamine M-1000 leads to water soluble and biodegradable [...] Read more.
A novel series of peptide based hybrid polymers designed to undergo enzymatic degradation is presented, via macrosubstitution of a polyphosphazene backbone with the tetrapeptide Gly-Phe-Leu-Gly. Further co-substitution of the hybrid polymers with hydrophilic polyalkylene oxide Jeffamine M-1000 leads to water soluble and biodegradable hybrid polymers. Detailed degradation studies, via 31P NMR spectroscopy, dynamic light scattering and field flow fractionation show the polymers degrade via a combination of enzymatic, as well as hydrolytic pathways. The peptide sequence was chosen due to its known property to undergo lysosomal degradation; hence, these degradable, water soluble polymers could be of significant interest for the use as polymer therapeutics. In this context, we investigated conjugation of the immune response modifier imiquimod to the polymers via the tetrapeptide and report the self-assembly behavior of the conjugate, as well as its enzymatically triggered drug release behavior. Full article
(This article belongs to the Special Issue Biodegradable Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Living polymerisation of poly(dichlorophosphazene) and macromolecular substitution of the chlorine atoms by Gly-Phe-Leu-Gly-OtBu.</p>
Full article ">Figure 2
<p>(<b>A</b>) Poly(organo)phosphazenes <b>2</b>–<b>4</b> with the tetrapeptide linker GFLG and Jeffamine M-1000 coupled to the polyphosphazene backbone. (<b>B</b>) Poly(organo)phosphazene <b>5</b> loaded with imiquimod via the tetrapeptide linker GFLG and Jeffamine M-1000 coupled via a glycine spacer to the polyphosphazene backbone. The combinations of the two different side chains are statistically distributed.</p>
Full article ">Figure 3
<p>Molecular size distribution by intensity (<b>A</b>) and volume (<b>B</b>) as detected by dynamic light scattering for polymers <b>2</b>–<b>5</b> in phosphate buffer at pH 7.4 (polymer concentration 1 mg/mL, d<sub>h</sub>—hydrodynamic diameter).</p>
Full article ">Figure 4
<p>Intra- and intermolecular self-assembly of poly(organo)phosphazenes with amphiphilic character in aqueous solution.</p>
Full article ">Figure 5
<p>Normalized FFF analysis illustrating the degradation of polymer <b>2</b> at 37 °C, in an aqueous solution at pH: 2 (<b>a</b>); 5 (<b>b</b>); and 7.4 (<b>c</b>). Broadening and decrease in intensity and a shift to earlier elution time of the polymer peak are observed.</p>
Full article ">Figure 6
<p>Size distribution by volume measured by dynamic light scattering for polymer <b>5</b> in citrate/phosphate buffer at pH 2 at various time points (polymer concentration—1 mg/mL, D<sub>h</sub>—hydrodynamic diameter).</p>
Full article ">Figure 7
<p>Enzymatic degradation of polymer <b>2</b> followed by <sup>31</sup>P NMR spectroscopy over 28 days in citrate buffer (pH6) containing L-cysteine and papain (<b>a</b>); hydrolytic degradation of polymer <b>2</b> in the same buffer system without papain (<b>b</b>); and with papain and cystamine as inhibitor (<b>c</b>). All samples were stored at 37 °C.</p>
Full article ">Figure 8
<p>Phosphate determination of polymer <b>2</b> quantitatively determined by UV–Vis analysis to show the degradation profile of the polymer in aqueous conditions at pH 5 (<math display="inline"> <mrow> <mo mathcolor="blue" mathsize="80%">■</mo> </mrow> </math>) and with papain at pH 5 (<math display="inline"> <mrow> <mo mathcolor="red">●</mo> </mrow> </math>)at 37 °C.</p>
Full article ">Figure 9
<p>Preferential cleavable site of papain and proposed hydrolytic and enzyme initiated degradation mechanism of GFLG-peptide based poly(organo)phosphazenes.</p>
Full article ">Figure 10
<p>Hydrolytic release of imiquimod from polymer <b>5</b> at 37 °C in acidic environment (citrate buffer, pH 6) (<math display="inline"> <mrow> <mo mathcolor="red">●</mo> </mrow> </math>), and enzymatic release of imiquimod from polymer <b>5</b> at 37 °C in the same buffer with L-cysteine activated papain (■). The amount of the released drug was estimated using a calibration curve for the free drug.</p>
Full article ">
2476 KiB  
Article
Synthesis of Highly Branched Polyolefins Using Phenyl Substituted α-Diimine Ni(II) Catalysts
by Fuzhou Wang, Ryo Tanaka, Zhengguo Cai, Yuushou Nakayama and Takeshi Shiono
Polymers 2016, 8(4), 160; https://doi.org/10.3390/polym8040160 - 22 Apr 2016
Cited by 39 | Viewed by 10024
Abstract
A series of α-diimine Ni(II) complexes containing bulky phenyl groups, [ArN = C(Naphth)C = NAr]NiBr2 (Naphth: 1,8-naphthdiyl, Ar = 2,6-Me2-4-PhC6H2 (C1); Ar = 2,4-Me2-6-PhC6H2 (C2); Ar = 2-Me-4,6-Ph2C6H [...] Read more.
A series of α-diimine Ni(II) complexes containing bulky phenyl groups, [ArN = C(Naphth)C = NAr]NiBr2 (Naphth: 1,8-naphthdiyl, Ar = 2,6-Me2-4-PhC6H2 (C1); Ar = 2,4-Me2-6-PhC6H2 (C2); Ar = 2-Me-4,6-Ph2C6H2 (C3); Ar = 4-Me-2,6-Ph2C6H2 (C4); Ar = 4-Me-2-PhC6H3 (C5); Ar = 2,4,6-Ph3C6H2 (C6)), were synthesized and characterized. Upon activation with either diethylaluminum chloride (Et2AlCl) or modified methylaluminoxane (MMAO), all Ni(II) complexes showed high activities in ethylene polymerization and produced highly branched amorphous polyethylene (up to 145 branches/1000 carbons). Interestingly, the sec-butyl branches were observed in polyethylene depending on polymerization temperature. Polymerization of 1-alkene (1-hexene, 1-octene, 1-decene and 1-hexadecene) with C1-MMAO at room temperature resulted in branched polyolefins with narrow Mw/Mn values (ca. 1.2), which suggested a living polymerization. The polymerization results indicated the possibility of precise microstructure control, depending on the polymerization temperature and types of monomers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structures of the catalyst precursor <b>C1</b> with 50% probability level, and H atoms have been omitted for clarity. Selected bond lengths (Å) and angles (deg) for complexes: Br1–Ni1, 2.3343(12); Ni1–N1, 2.041(5); N1–C9, 1.290(9); N1–C8, 1.427(8); Br1–Ni1–Br1i, 117.34(8); N1–Ni1–Br1i, 112.96(6); C9–N1–C8, 118.0(6).</p>
Full article ">Figure 2
<p><sup>13</sup>C NMR spectrum of polyethylenes obtained with <b>C1</b>-MMAO at 0–80 °C (A and B refer to methyl carbon of <span class="html-italic">sec-</span>butyl branches, entries 10–14, <a href="#polymers-08-00160-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 3
<p>Plots of <span class="html-italic">M</span><sub>n</sub> (■) and <span class="html-italic">M</span><sub>w</sub>/<span class="html-italic">M</span><sub>n</sub> (▲) as a function of reaction time for the polymerization of 1-decene (20 °C, <a href="#app1-polymers-08-00160" class="html-app">Table S3</a>).</p>
Full article ">Figure 4
<p><sup>13</sup>C NMR spectra of poly(1-decene)s obtained with <b>C1</b>-MMAO at −20–80 °C (entries 6–10, <a href="#polymers-08-00160-t003" class="html-table">Table 3</a>).</p>
Full article ">Figure 5
<p><sup>13</sup>C NMR spectra of poly(1-hexadecene)s obtained with <b>C1</b>-MMAO at 0 and 20 °C (entries 13 and 14, <a href="#polymers-08-00160-t003" class="html-table">Table 3</a>).</p>
Full article ">Scheme 1
<p>Mechanism for ethylene polymerization and 1-alkenes enchainment (L: NN = diimine Ligand; <span class="html-italic">ω-</span>position, where <span class="html-italic">ω</span> is the number of carbon atoms).</p>
Full article ">Scheme 2
<p>Synthesis of α-diimine ligands <b>L1</b>–<b>L7</b> and their Ni(II) complexes<b> C1</b>–<b>C7</b>.</p>
Full article ">
2073 KiB  
Article
The Effects of in Situ-Formed Silver Nanoparticles on the Electrical Properties of Epoxy Resin Filled with Silver Nanowires
by Gwang-Seok Song, Dai Soo Lee and Ilho Kang
Polymers 2016, 8(4), 157; https://doi.org/10.3390/polym8040157 - 21 Apr 2016
Cited by 10 | Viewed by 6787
Abstract
A novel method for preparing epoxy/silver nanocomposites was developed via the in situ formation of silver nanoparticles (AgNPs) within the epoxy resin matrix while using silver nanowires (AgNWs) as a conductive filler. The silver–imidazole complex was synthesized from silver acetate (AgAc) and 1-(2-cyanoethyl)-2-ethyl-4-methylimidazole [...] Read more.
A novel method for preparing epoxy/silver nanocomposites was developed via the in situ formation of silver nanoparticles (AgNPs) within the epoxy resin matrix while using silver nanowires (AgNWs) as a conductive filler. The silver–imidazole complex was synthesized from silver acetate (AgAc) and 1-(2-cyanoethyl)-2-ethyl-4-methylimidazole (imidazole). AgNPs were generated in situ during the curing of the epoxy resin through the thermal decomposition of the AgAc–imidazole complex, which was capable of reducing Ag+ to Ag by itself. The released imidazole acted as a catalyst to cure the epoxy. Additionally, after the curing process, the in situ-generated AgNPs were stabilized by the formed epoxy network. Therefore, by using the thermal decomposition method, uniformly dispersed AgNPs of approximately 100 nm were formed in situ in the epoxy matrix filled with AgNWs. It was observed that the nanocomposites containing in situ-formed AgNPs exhibited isotropic electrical properties in the epoxy resins in the presence of AgNWs. Full article
(This article belongs to the Special Issue Nano- and Microcomposites for Electrical Engineering Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesized silver nanowires (AgNWs): (<b>a</b>) transmission electron microscopy image; (<b>b</b>) diameter distribution of the purified AgNW.</p>
Full article ">Figure 2
<p>Purified silver nanowires (AgNWs): (<b>a</b>) scanning electron microscopy image; (<b>b</b>) length distribution of a purified AgNW.</p>
Full article ">Figure 3
<p>Thermogravimetric analysis thermograms of imidazole, silver–imidazole complex (AgI), and silver acetate (AgAc).</p>
Full article ">Figure 4
<p>Surface electrical resistivity of the cured epoxy resin containing AgNWs of various concentrations: (<b>a</b>) sample cured with imidazole in the direction of the highest resistivity; (<b>b</b>) sample cured with imidazole in the direction perpendicular to that of (<b>a</b>); (<b>c</b>) sample cured with silver–imidazole complex in the direction of the highest resistivity; (<b>d</b>) sample cured with silver–imidazole complex in the direction perpendicular to that of (<b>c</b>).</p>
Full article ">Figure 5
<p>Interwire distances of AgNWs of different diameters: (<b>a</b>) 10 nm; (<b>b</b>) 50 nm; (<b>c</b>) 100 nm; (<b>d</b>) 200 nm; and (<b>e</b>) 500 nm.</p>
Full article ">Figure 6
<p>TEM images of the cured epoxy resin/silver nanocomposites: (<b>a</b>) sample cured with imidazole (scale bar, 200 nm); (<b>b</b>) sample cured with 10.0 wt % of the silver–imidazole complex with 6 vol % of AgNWs (scale bar, 1 μm); (<b>c</b>) magnified version of the sample in (<b>b</b>) (scale bar, 200 nm); (<b>d</b>) sample cured with 10.0 wt % of silver–imidazole with 8 vol % of AgNWs (scale bar, 200 nm).</p>
Full article ">Scheme 1
<p>Scheme for the silver nanoparticles (AgNPs) formed by <span class="html-italic">in situ</span> thermal decomposition of silver–imidazole complex.</p>
Full article ">
7028 KiB  
Review
Design and Utility of Metal/Metal Oxide Nanoparticles Mediated by Thioether End-Functionalized Polymeric Ligands
by Shumaila Razzaque, Syed Zajif Hussain, Irshad Hussain and Bien Tan
Polymers 2016, 8(4), 156; https://doi.org/10.3390/polym8040156 - 21 Apr 2016
Cited by 48 | Viewed by 12756
Abstract
The past few decades have witnessed significant advances in the development of functionalized metal/metal oxide nanoparticles including those of inorganic noble metals and magnetic materials stabilized by various polymeric ligands. Recent applications of such functionalized nanoparticles, including those in bio-imaging, sensing, catalysis, drug [...] Read more.
The past few decades have witnessed significant advances in the development of functionalized metal/metal oxide nanoparticles including those of inorganic noble metals and magnetic materials stabilized by various polymeric ligands. Recent applications of such functionalized nanoparticles, including those in bio-imaging, sensing, catalysis, drug delivery, and other biomedical applications have triggered the need for their facile and reproducible preparation with a better control over their size, shape, and surface chemistry. In this perspective, the multidentate polymer ligands containing functional groups like thiol, thioether, and ester are important surface ligands for designing and synthesizing stable nanoparticles (NPs) of metals or their oxides with reproducibility and high yield. These ligands have offered an unprecedented control over the particle size of both nanoparticles and nanoclusters with enhanced colloidal stability, having tunable solubility in aqueous and organic media, and tunable optical, magnetic, and fluorescent properties. This review summarizes the synthetic methodologies and stability of nanoparticles and fluorescent nanoclusters of metals (Au, Ag, Cu, Pt, and other transition metal oxides) prepared by using thioether based ligands and highlights their applications in bio-imaging, sensing, drug delivery, magnetic resonance imaging (MRI), and catalysis. The future applications of fluorescent metal NPs like thermal gradient optical imaging, single molecule optoelectronics, sensors, and optical components of the detector are also envisaged. Full article
(This article belongs to the Collection Featured Mini Reviews in Polymer Science)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of (<b>a</b>) direct synthesis (in situ) of nanoparticles (NPs) (<b>b</b>) Post synthesis capping by ligand exchange method for the polymer stabilized NPs.</p>
Full article ">Figure 2
<p>Transmission electron microscopy (TEM) micrographs of Gold NPs with various concentrations of polymer ligands. (<b>A</b>) 0.006mM (<b>B</b>) 0.03mM (<b>C</b>) 0.6mM and (<b>D</b>) 3.6mM [<a href="#B41-polymers-08-00156" class="html-bibr">41</a>].</p>
Full article ">Figure 3
<p>TEM micrographs and the corresponding size distribution histograms of the Co NPs synthesized by rapid injection (<b>top</b> panels) or drop-wise addition (<b>bottom</b> panels) of the reductant NaBH<sub>4</sub> into a mixture of CoCl<sub>2</sub> solution and the polymer [<a href="#B62-polymers-08-00156" class="html-bibr">62</a>].</p>
Full article ">Figure 4
<p>Scheme illustrating (<b>a</b>) The preparation of AuNCs in organic solvent; (<b>b</b>) TEM images of AuNCs capped by thioether end-functionalized ligands. The scale bar is 20 nm.</p>
Full article ">Figure 5
<p>Scheme illustrating the (<b>i</b>) Synthesis of MIONs; (<b>ii</b>) TEM images (<b>a,b</b>) 4.5-nm and (<b>c,d</b>) 8.5-nm PMAA–PTTM protected Fe<sub>3</sub>O<sub>4</sub> nanocrystals. The molar ratio between carboxylic acid groups and FeCl<sub>3</sub>·6H<sub>2</sub>O in both cases is 3:4 [<a href="#B57-polymers-08-00156" class="html-bibr">57</a>].</p>
Full article ">Figure 6
<p>(<b>A</b>) Structural presentation of MIONs@DDT–PMAA. (<b>B</b>) and (<b>C</b>) TEM images and DLS curves along with particle size distribution histograms of MIONs prepared with (<b>a</b>) 1.5 mM of 0.5% DDT–PMAA (<b>b</b>); 2% DDT–PMAA (<b>c</b>); 5% DDT–PMAA (<b>d</b>) and 10% DDT–PMAA [<a href="#B7-polymers-08-00156" class="html-bibr">7</a>].</p>
Full article ">Figure 6 Cont.
<p>(<b>A</b>) Structural presentation of MIONs@DDT–PMAA. (<b>B</b>) and (<b>C</b>) TEM images and DLS curves along with particle size distribution histograms of MIONs prepared with (<b>a</b>) 1.5 mM of 0.5% DDT–PMAA (<b>b</b>); 2% DDT–PMAA (<b>c</b>); 5% DDT–PMAA (<b>d</b>) and 10% DDT–PMAA [<a href="#B7-polymers-08-00156" class="html-bibr">7</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Reaction scheme for the synthesis of thioether polymer ligand PTMP–PMAA; (<b>b</b>) Preparation of Au NCs stabilized by polymer ligands. Inserts are photographs of Au NCs’ aqueous solution under the irradiation of 365 nm ultraviolet light (<b>left, red</b> fluorescence) and day light (<b>right</b>, <b>yellow</b> color) [<a href="#B94-polymers-08-00156" class="html-bibr">94</a>].</p>
Full article ">Figure 8
<p>Schematic diagram of preparation of (<b>a</b>) the polymer ligand PTMP–PMAA; (<b>b</b>) photoreductive synthesis of fluorescent Cu, Ag, and Au nanoclusters; (<b>c</b>) TEM image of Au nanoclusters; and (<b>d</b>) Time dependent evolution of fluorescence emission spectra of the solution containing (<b>i</b>) PTMP–PMAA and Cu(NO<sub>3</sub>)<sub>2</sub> (excited at 360 nm); (<b>ii</b>) PTMP–PMAA and AgNO<sub>3</sub> (excited at 405 nm); (<b>iii</b>) PTMP–PMAA and HAuCl<sub>4</sub> (excited at 360 nm) upon UV-irradiation at 365 nm [<a href="#B5-polymers-08-00156" class="html-bibr">5</a>].</p>
Full article ">Figure 9
<p>Confocal microscopic Images of (<b>a</b>) CBMC; normal cells; (<b>b</b>) K562; cancer cells, after incubation with Au NCs in a medium containing fetal calf serum (FCS) for 24 h. The nuclei of cells were tainted with Hochest-33258 to yield blue fluorescence and the red fluorescence is due to the presence of Au NCs [<a href="#B94-polymers-08-00156" class="html-bibr">94</a>]. The scale bar is 20 µm.</p>
Full article ">Figure 10
<p>Viability of cells (HepG2) after (<b>a</b>) 24 h; (<b>b</b>) 48 h; and (<b>c</b>) 72 h incubation with DOX–MIONs, DOX/MIONs, and free DOX having equivalent concentration of DOX [<a href="#B7-polymers-08-00156" class="html-bibr">7</a>].</p>
Full article ">Figure 11
<p>Formation of Au species during the one-pot synthesis process: (<b>I</b>) mixing of all reaction precursors in one pot; (<b>II</b>) hydrothermal treatment at 100 °C for 24 h; (<b>III</b>) ethanol extraction followed by H<sub>2</sub> reduction; and (<b>III′</b>) calcined at high-temperature [<a href="#B133-polymers-08-00156" class="html-bibr">133</a>].</p>
Full article ">Scheme 1
<p>Schematic illustration for the preparation of Au nanoparticles (NPs) stabilized by thioether polymer ligands [<a href="#B29-polymers-08-00156" class="html-bibr">29</a>].</p>
Full article ">Scheme 2
<p>Schematic illustration of thioether polymer ligand formation [<a href="#B29-polymers-08-00156" class="html-bibr">29</a>].</p>
Full article ">
2456 KiB  
Article
Effect of Small Reaction Locus in Free-Radical Polymerization: Conventional and Reversible-Deactivation Radical Polymerization
by Hidetaka Tobita
Polymers 2016, 8(4), 155; https://doi.org/10.3390/polym8040155 - 20 Apr 2016
Cited by 8 | Viewed by 5433
Abstract
When the size of a polymerization locus is smaller than a few hundred nanometers, such as in miniemulsion polymerization, each locus may contain no more than one key-component molecule, and the concentration may become much larger than the corresponding bulk polymerization, leading to [...] Read more.
When the size of a polymerization locus is smaller than a few hundred nanometers, such as in miniemulsion polymerization, each locus may contain no more than one key-component molecule, and the concentration may become much larger than the corresponding bulk polymerization, leading to a significantly different rate of polymerization. By focusing attention on the component having the lowest concentration within the species involved in the polymerization rate expression, a simple formula can predict the particle diameter below which the polymerization rate changes significantly from the bulk polymerization. The key component in the conventional free-radical polymerization is the active radical and the polymerization rate becomes larger than the corresponding bulk polymerization when the particle size is smaller than the predicted diameter. The key component in reversible-addition-fragmentation chain-transfer (RAFT) polymerization is the intermediate species, and it can be used to predict the particle diameter below which the polymerization rate starts to increase. On the other hand, the key component is the trapping agent in stable-radical-mediated polymerization (SRMP) and atom-transfer radical polymerization (ATRP), and the polymerization rate decreases as the particle size becomes smaller than the predicted diameter. Full article
(This article belongs to the Special Issue Selected Papers from ASEPFPM2015)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the radical concentration in (<b>a</b>) bulk; (<b>b</b>) pseudo-bulk; and (<b>c</b>) miniemulsion polymerization.</p>
Full article ">Figure 2
<p>Calculated conversion development for bulk and miniemulsion polymerization with <span class="html-italic">R</span><sub>I</sub> = 1 × 10<sup>−7</sup> mol·L<sup>−1</sup>·s<sup>−1</sup>, <span class="html-italic">k</span><sub>t</sub> = 1 × 10<sup>7</sup> L·mol<sup>−1</sup>·s<sup>−1</sup>, <math display="inline"> <semantics> <mover accent="true"> <mi>n</mi> <mo>¯</mo> </mover> </semantics> </math> = 0.5 and <span class="html-italic">k</span><sub>p</sub> = 500 L·mol<sup>−1</sup>·s<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Reversible deactivation reaction scheme in each type of RDRP. In the figure, P<span class="html-italic"><sub>i</sub></span>X or XP<span class="html-italic"><sub>i</sub></span> is the dormant polymer with chain length <span class="html-italic">i</span>. <math display="inline"> <semantics> <mrow> <msubsup> <mi mathvariant="normal">R</mi> <mi>i</mi> <mo>•</mo> </msubsup> </mrow> </semantics> </math> is the active polymer radical with chain length <span class="html-italic">i</span>.</p>
Full article ">Figure 4
<p>Calculated polymerization rate for the TEMPO-mediated styrene polymerization at 10% conversion with the initial RGS concentration, [RGS]<sub>0</sub> = 0.2 mol/L. The data (symbols) are taken from [<a href="#B14-polymers-08-00155" class="html-bibr">14</a>].</p>
Full article ">Figure 5
<p>Schematic representation of the zero-one behavior in the conventional free-radical polymerization (<b>red</b>) and RAFT (<b>blue</b>) miniemulsion polymerization, where <span class="html-italic">n</span> is the number of radicals in a particle in the conventional free-radical polymerization and <span class="html-italic">n</span><sub>RGS</sub> is the number of intermediate molecules in RAFT.</p>
Full article ">Figure 6
<p>Calculated threshold diameter change during RAFT polymerization. The parameters used are: <span class="html-italic">R</span><sub>I</sub> = 1 × 10<sup>−7</sup> mol·L<sup>−1</sup>·s<sup>−1</sup>, <span class="html-italic">k</span><sub>p</sub> = 500 L·mol<sup>−1</sup>·s<sup>−1</sup>, <span class="html-italic">k</span><sub>t</sub> = 1 × 10<sup>7</sup> L·mol<sup>−1</sup>·s<sup>−1</sup>, [M]<sub>0</sub> = 8 mol·L<sup>−1</sup> and <span class="html-italic">k</span><sub>2</sub> = 1 × 10<sup>6</sup> L·mol<sup>−1</sup>·s<sup>−1</sup> for both models. For the IT model, <span class="html-italic">k</span><sub>1</sub> = 1 × 10<sup>4</sup> s<sup>−1</sup> and <span class="html-italic">k</span><sub>t,RGS</sub> = 1×10<sup>7</sup> L·mol<sup>−1</sup>·s<sup>−1</sup>. For the SF model, <span class="html-italic">k</span><sub>1</sub> = 0.5 s<sup>−1</sup> and <span class="html-italic">k</span><sub>t,RGS</sub> = 0.</p>
Full article ">Figure 7
<p>Monte Carlo simulation results for IT and SF model, with the same set of parameters used in <a href="#polymers-08-00155-f006" class="html-fig">Figure 6</a>, based on (<b>a</b>) the IT model and (<b>b</b>) the SF model. Reproduced from <a href="#polymers-08-00155-f001" class="html-fig">Figure 1</a> in [<a href="#B23-polymers-08-00155" class="html-bibr">23</a>]. © Copyright permission from Wiley-VCH Verlag GmbH &amp; Co. KGaA.</p>
Full article ">Figure 8
<p>Conversion development during polystyryl dithiobenzoate-mediated styrene polymerization at 60 °C [<a href="#B25-polymers-08-00155" class="html-bibr">25</a>]. The solid curves are the calculation results and symbols are the experimental results. For the calculation, the differential equations given by Equations (31)–(34) are solved for the bulk polymerization, and for miniemulsion polymerization, the MC simulation was employed for each fixed diameter. Reproduced from the graphical abstract of [<a href="#B25-polymers-08-00155" class="html-bibr">25</a>]. © Copyright permission from Wiley-VCH Verlag GmbH &amp; Co. KGaA.</p>
Full article ">
3008 KiB  
Article
Poly-Lactide/Exfoliated C30B Interactions and Influence on Thermo-Mechanical Properties Due to Artificial Weathering
by Wendy Margarita Chávez-Montes, Guillermo González-Sánchez and Sergio Gabriel Flores-Gallardo
Polymers 2016, 8(4), 154; https://doi.org/10.3390/polym8040154 - 20 Apr 2016
Cited by 6 | Viewed by 7288
Abstract
Thermal stability as well as enhanced mechanical properties of poly-lactide (PLA) can increase PLA applications for short-use products. The conjunction of adequate molecular weight (MW) as well as satisfactory thermo-mechanical properties, together, can lead to the achievement of suitable properties. [...] Read more.
Thermal stability as well as enhanced mechanical properties of poly-lactide (PLA) can increase PLA applications for short-use products. The conjunction of adequate molecular weight (MW) as well as satisfactory thermo-mechanical properties, together, can lead to the achievement of suitable properties. However, PLA is susceptible to thermal degradation and thus an undesired decay of MW and a decrease of its mechanical properties during processing. To avoid this PLA degradation, nanofiller is incorporated as reinforcement to increase its thermo-mechanical properties. There are many papers focusing on filler effects on the thermal stability and mechanical properties of PLA/nanocomposites; however, these investigations lack an explanation of polymer/filler interactions. We propose interactions between PLA and Cloisite30B (C30B) as nanofiller. We also study the effects on the thermal and mechanical properties due to molecular weight decay after exposure to artificial weathering. PLA blank and nanocomposites were subjected to three time treatments (0, 176, and 360 h) of exposure to artificial weathering in order to achieve comparable materials with different MW. MW was acquired by means of Gel Permeation Chromatography (GPC). Thermo-mechanical properties were investigated through Thermogravimetric Analysis (TGA), Differential Scanning Calorimetry (DSC), X-ray Diffraction (XRD), Dynamic Mechanical Thermal Analysis (DMTA) and Fourier Transform Infrared Spectroscopy (FTIR). Full article
(This article belongs to the Special Issue Biodegradable Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Thermogravimetric analysis of C30B exposed to both thermal treatment and artificial weathering.</p>
Full article ">Figure 2
<p>Thermogravimetric analysis (TGA) of (<b>a</b>) PLA/B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of A = 49.0, B = 33.1, and C = 16.4) and; (<b>b</b>) PLA/C30B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of D = 43.5, E = 31.9 and F = 11.3).</p>
Full article ">Figure 3
<p>Differential Scanning Calorimetry (DSC) curves of PLA/B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of A = 49.0, B = 33.1, and C = 16.4) and PLA/C30B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of D = 43.5, E = 31.9 and F = 11.3).</p>
Full article ">Figure 4
<p>X-ray diffraction (XRD) diffractograms of (<b>a</b>) PLA/B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of A = 49.0, B = 33.1, and C = 16.4); and (<b>b</b>) PLA/C30B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of D = 43.5, E = 31.9, and F = 11.3).</p>
Full article ">Figure 5
<p>Results of Dynamic Thermo-Mechanical Analysis (DTMA): Storage modulus (<span class="html-italic">E</span>′) and tan δ of (<b>a</b>) PLA/B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of A = 49.0, B = 33.1, and C = 16.4) and; (<b>b</b>) PLA/C30B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of D = 43.5, E = 31.9, and F = 11.3).</p>
Full article ">Figure 6
<p>Suggested interactions between the C30B organomodifier and PLA.</p>
Full article ">Figure 7
<p>IR spectra of (<b>a</b>) PLA/B and PLA/C30B (0 h of exposure to artificial weathering) and; (<b>b</b>) PLA/B and PLA/C30B (<span class="html-italic">M</span><sub>n</sub> in kg·mol<sup>−1</sup> of D = 43.5, E = 31.9, and F = 11.3).</p>
Full article ">
1882 KiB  
Article
Physical and Mechanical Evaluation of Five Suture Materials on Three Knot Configurations: An in Vitro Study
by Desire Abellán, José Nart, Andrés Pascual, Robert E. Cohen and Javier D. Sanz-Moliner
Polymers 2016, 8(4), 147; https://doi.org/10.3390/polym8040147 - 20 Apr 2016
Cited by 36 | Viewed by 9971
Abstract
The aim of this study was to evaluate and compare the mechanical properties of five suture materials on three knot configurations when subjected to different physical conditions. Five 5-0 (silk, polyamide 6/66, polyglycolic acid, glycolide-e-caprolactone copolymer, polytetrafluoroethylene) suture materials were used. Ten samples [...] Read more.
The aim of this study was to evaluate and compare the mechanical properties of five suture materials on three knot configurations when subjected to different physical conditions. Five 5-0 (silk, polyamide 6/66, polyglycolic acid, glycolide-e-caprolactone copolymer, polytetrafluoroethylene) suture materials were used. Ten samples per group of each material were used. Three knot configurations were compared A.2=1=1 (forward–forward–reverse), B.2=1=1 (forward–reverse–forward), C.1=2=1 (forward–forward–reverse). Mechanical properties (failure load, elongation, knot slippage/breakage) were measured using a universal testing machine. Samples were immersed in three different pH concentrations (4,7,9) at room temperature for 7 and 14 days. For the thermal cycle process, sutures were immersed in two water tanks at different temperatures (5 and 55 °C). Elongation and failure load were directly dependent on the suture material. Polyglycolic acid followed by glycolide-e-caprolactone copolymer showed the most knot failure load, while polytetrafluoroethylene showed the lowest (P < 0.001). Physical conditions had no effect on knot failure load (P = 0.494). Statistically significant differences were observed between knot configurations (P = 0.008). Additionally, individual assessment of suture material showed statistically significant results for combinations of particular knot configurations. Physical conditions, such as pH concentration and thermal cycle process, have no influence on suture mechanical properties. However, knot failure load depends on the suture material and knot configuration used. Consequently, specific suturing protocols might be recommended to obtain higher results of knot security. Full article
(This article belongs to the Special Issue Biodegradable Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sample preparation. (<b>a</b>) Suture material before tying; (<b>b</b>) Tying of suture material around a 26 mm metal cylinder; (<b>c</b>) Different suture materials after tying; (<b>d</b>) Sample of suture material prepared for testing mechanical and chemical conditions.</p>
Full article ">Figure 2
<p>Mechanical and physical tests. (<b>a</b>) Hydraulic grip of the Universal Testing Machine; (<b>b</b>) 5 and 55 °C water tanks of the Universal Thermal Cycling Testing Machine; (<b>c</b>) Metal boxes which contained the sample for the thermal cycle process.</p>
Full article ">Figure 3
<p>Tables for variables elongation and load failure (<b>a</b>) Elongation of suture materials at the three knot configurations; (<b>b</b>) Load failure of the suture materials to the different knot configurations.</p>
Full article ">Figure 4
<p>Comparison of suture material failure load in three knot configurations.</p>
Full article ">Figure 5
<p>Knot configurations failure load according to the suture material used.</p>
Full article ">Figure 6
<p>(<b>a</b>) Failure load of knots according to each suture material. (<b>b</b>) Frequency of knot breakage in relation to suture material and knot configuration used.</p>
Full article ">
4080 KiB  
Article
Bio-Based Resin Reinforced with Flax Fiber as Thermorheologically Complex Materials
by Ali Amiri, Arvin Yu, Dean Webster and Chad Ulven
Polymers 2016, 8(4), 153; https://doi.org/10.3390/polym8040153 - 19 Apr 2016
Cited by 30 | Viewed by 7946
Abstract
With the increase in structural applications of bio-based composites, the study of long-term creep behavior of these materials turns into a significant issue. Because of their bond type and structure, natural fibers and thermoset resins exhibit nonlinear viscoelastic behavior. Time-temperature superposition (TTS) provides [...] Read more.
With the increase in structural applications of bio-based composites, the study of long-term creep behavior of these materials turns into a significant issue. Because of their bond type and structure, natural fibers and thermoset resins exhibit nonlinear viscoelastic behavior. Time-temperature superposition (TTS) provides a useful tool to overcome the challenge of the long time required to perform the tests. The TTS principle assumes that the effect of temperature and time are equivalent when considering the creep behavior, therefore creep tests performed at elevated temperatures may be converted to tests performed at longer times. In this study, flax fiber composites were processed with a novel liquid molding methacrylated epoxidized sucrose soyate (MESS) resin. Frequency scans of flax/MESS composites were obtained at different temperatures and storage modulus and loss modulus were recorded and the application of horizontal and vertical shift factors to these viscoelastic functions were studied. In addition, short-term strain creep at different temperatures was measured and curves were shifted with solely horizontal, and with both horizontal and vertical shift factors. The resulting master curves were compared with a 24-h creep test and two extrapolated creep models. The findings revealed that use of both horizontal and vertical shift factors will result in a smoother master curves for loss modulus and storage modulus, while use of only horizontal shift factors for creep data provides acceptable creep strain master curves. Based on the findings of this study, flax/MESS composites can be considered as thermorheologically complex materials. Full article
(This article belongs to the Special Issue Renewable Polymeric Adhesives)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images of (<b>a</b>) the cross-section of flax fiber and (<b>b</b>) the surface of the fiber used in this study.</p>
Full article ">Figure 2
<p>Synthetic route to MESS.</p>
Full article ">Figure 3
<p>Schematic of the Vacuum Assisted Resin Transfer Molding (VARTM) set-up used to manufacture composite sample.</p>
Full article ">Figure 4
<p>Fourier transform infrared spectrum of MESS.</p>
Full article ">Figure 5
<p>Proton nuclear magnetic resonance spectrum of the methacrylated epoxidized sucrose soyate (MESS) in CDCl<sub>3</sub>.</p>
Full article ">Figure 6
<p>Frequency sweep of flax/MESS composite at different temperatures.</p>
Full article ">Figure 7
<p>Master curves generated by solely horizontal shifting of storage modulus curve and using the same shift factors for loss modulus curves.</p>
Full article ">Figure 8
<p>Master curves obtained by horizontal and vertical shifting of the frequency sweeps.</p>
Full article ">Figure 9
<p>(<b>a</b>) Horizontal shift factors when only horizontal shift factors are used; (<b>b</b>) horizontal and vertical shift factors when both are used.</p>
Full article ">Figure 10
<p>Creep strain <span class="html-italic">vs.</span> time at different temperatures.</p>
Full article ">Figure 11
<p>Creep strain master curve at 30 °C obtained by horizontal shifting of creep data at different temperatures.</p>
Full article ">Figure 12
<p>Creep strain curves at different temperatures shifted by the horizontal shift factors obtained from storage modulus master curve.</p>
Full article ">Figure 13
<p>Creep strain master curve generated by horizontal and vertical shift factors.</p>
Full article ">Figure 14
<p>Comparison of extrapolated creep data with Nutting and Findley Power Laws with actual creep data for 24 h.</p>
Full article ">Figure 15
<p>Comparison of actual creep data for 24 h with (<b>a</b>) master curve generated by horizontal shifting of creep data; (<b>b</b>) master curve generated by horizontal and vertical shift of creep data.</p>
Full article ">
1746 KiB  
Article
Opening New Gates for the Modification of PVC or Other PVC Derivatives: Synthetic Strategies for the Covalent Binding of Molecules to PVC
by Rodrigo Navarro, Mónica Pérez Perrino, Carolina García, Carlos Elvira, Alberto Gallardo and Helmut Reinecke
Polymers 2016, 8(4), 152; https://doi.org/10.3390/polym8040152 - 19 Apr 2016
Cited by 36 | Viewed by 9519
Abstract
Several synthetic strategies based on the use of substituted aromatic and hetero-aromatic thiols for the covalent binding of modifier compounds to PVC are described. A variety of aliphatic alcohols and amines are linked to the aromatic or heteroaromatic rings via highly active functionalities [...] Read more.
Several synthetic strategies based on the use of substituted aromatic and hetero-aromatic thiols for the covalent binding of modifier compounds to PVC are described. A variety of aliphatic alcohols and amines are linked to the aromatic or heteroaromatic rings via highly active functionalities as the isocyanate, acidchloride, or chlorosulfonyl group, and the three chlorine atoms of trichlorotriazine. The first three pathways lead to protected aromatic disulfides obtaining the substituted aromatic thiols by reduction as a final step of an unprecedented synthetic route. The second approach, in a novel, extremely efficient, and scalable process, uses the particular selectivity of trichlorotriazine to connect aliphatic amines, alcohols, and thiols to the ring and creates the thiol via nucleophilic substitution of a heteroaromatic halogen by thiourea and subsequent hydrolysis. Most of the modifier compounds were linked to the polymer chains with high degrees of anchorage. The presented approaches are highly versatile as different activations of aromatic and heteroaromatic rings are used. Therefore, many types of tailored functional nucleophiles may be anchored to PVC providing non-migrating materials with a broad range of applications and properties. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Series of <sup>1</sup>H–NMR spectra of PVC modified with (<b>a</b>) 20 wt % and (<b>b</b>) 40 wt % of compound <b>oE</b>; and (<b>c</b>) shows the spectrum of pure <b>oE</b>.</p>
Full article ">Figure 2
<p>IR spectra of TCTA-derivatives of modifier <b>6</b> (see <a href="#polymers-08-00152-t002" class="html-table">Table 2</a>) being R = –N(C<sub>8</sub>H<sub>17</sub>)<sub>2</sub>.</p>
Full article ">Figure 3
<p><sup>13</sup>C–NMR spectra of derivatives of modifier <b>6</b>: (<b>a</b>) TCTA-Cl; (<b>b</b>) TCTA-thiouronium salt; (<b>c</b>) TCTA-SH and (<b>d</b>) TCTA-SNa is 165.6 ppm. Similarly, the C–S shift of the thiouronium salt.</p>
Full article ">Figure 4
<p><sup>1</sup>H-NMR spectra of PVC modified with different amounts of compound <b>7</b>. (<b>a</b>) 6 mol %; (<b>b</b>) 11 mol %; (<b>c</b>) 22 mol%; (<b>d</b>) 34 mol%, and (<b>e</b>) pure compound <b>7</b>.</p>
Full article ">Scheme 1
<p>Synthesis of novel PVC-linkable modifiers.</p>
Full article ">Scheme 2
<p>Structures of novel PVC-linkable modifiers.</p>
Full article ">Scheme 3
<p>Synthesis of TCTA based PVC-linkable compounds. The structures of <span class="html-italic">Y</span> and <span class="html-italic">R</span> can be found in <a href="#polymers-08-00152-t002" class="html-table">Table 2</a>.</p>
Full article ">
10111 KiB  
Article
The in Vitro and in Vivo Degradation of Cross-Linked Poly(trimethylene carbonate)-Based Networks
by Liqun Yang, Jianxin Li, Miao Li and Zhongwei Gu
Polymers 2016, 8(4), 151; https://doi.org/10.3390/polym8040151 - 19 Apr 2016
Cited by 31 | Viewed by 8212
Abstract
The degradation of the poly(trimethylene carbonate) (PTMC) and poly(trimethylene carbonate-co-ε-caprolactone) (P(TMC-co-CL)) networks cross-linked by 0.01 and 0.02 mol % 2,2′-bis(trimethylene carbonate-5-yl)-butylether (BTB) was carried out in the conditions of hydrolysis and enzymes in vitro and subcutaneous implantation in vivo [...] Read more.
The degradation of the poly(trimethylene carbonate) (PTMC) and poly(trimethylene carbonate-co-ε-caprolactone) (P(TMC-co-CL)) networks cross-linked by 0.01 and 0.02 mol % 2,2′-bis(trimethylene carbonate-5-yl)-butylether (BTB) was carried out in the conditions of hydrolysis and enzymes in vitro and subcutaneous implantation in vivo. The results showed that the cross-linked PTMC networks exhibited much faster degradation in enzymatic conditions in vitro and in vivo versus in a hydrolysis case due to the catalyst effect of enzymes; the weight loss and physical properties of the degraded networks were dependent on the BTB amount. The morphology observation in lipase and in vivo illustrated that enzymes played an important role in the surface erosion of cross-linked PTMC. The hydrolytic degradation rate of the cross-linked P(TMC-co-CL) networks increased with increasing ε-caprolactone (CL) content in composition due to the preferential cleavage of ester bonds. Cross-linking is an effective strategy to lower the degradation rate and enhance the form-stability of PTMC-based materials. Full article
(This article belongs to the Special Issue Functional Polymers for Medical Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the formation of the P(TMC-<span class="html-italic">co</span>-CL) networks cross-linked by BTB.</p>
Full article ">Figure 2
<p>Mass loss of N100 and N100-1 specimens conditioned in lipase solutions at 37 °C for different time periods.</p>
Full article ">Figure 3
<p>Shape of N100 (<b>A</b>) and N100-1 specimens (<b>B</b>) at different times of enzymatic degradation in lipase solutions. The initial diameter of the specimens was 10 mm.</p>
Full article ">Figure 4
<p>Relative thickness of N100 and N100-1 conditioned in lipase solutions at 37 °C for different times.</p>
Full article ">Figure 5
<p>SEM micrographs of N100-1 before and after 2, 10 and 15 week enzymatic degradation. The scale bar was 20 μm.</p>
Full article ">Figure 6
<p>Water uptake of N100 and N100-1 specimens conditioned in lipase solutions at 37 °C for different time periods.</p>
Full article ">Figure 7
<p>DSC (<b>A</b>) and TGA (<b>B</b>) curves of N100-1 conditioned in lipase solutions at 37 °C for different times.</p>
Full article ">Figure 8
<p>Change in thermal properties of N100-1 during <span class="html-italic">in vitro</span> enzymatic degradation in lipase solutions at 37 °C.</p>
Full article ">Figure 9
<p>Mass loss of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks during hydrolytic degradation.</p>
Full article ">Figure 10
<p>Macroscopic observation of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks during hydrolytic degradation: (<b>A</b>) N100-1; (<b>B</b>) N75-1; and (<b>C</b>) N50-1.</p>
Full article ">Figure 11
<p>SEM micrographs of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks before and after the <span class="html-italic">in vitro</span> hydrolytic degradation: (<b>A</b>) N100-1; (<b>B</b>) N75-1; (<b>C</b>) N50-1. The scale bar was 100 μm for all images with exception of 50 μm for N75-1 at week 30.</p>
Full article ">Figure 12
<p>Water uptake s of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks during hydrolytic degradation.</p>
Full article ">Figure 13
<p>Change in thermal properties of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks during <span class="html-italic">in vitro</span> hydrolytic degradation in PBS at 37 °C: (A) <span class="html-italic">T</span>g and (B) <span class="html-italic">T</span>d.</p>
Full article ">Figure 14
<p>Change in physical properties of the cross-linked P(TMC-<span class="html-italic">co</span>-CL) networks during <span class="html-italic">in vitro</span> degradation in pH 7.4 PBS: (<b>A</b>) Young’s modulus (<b>B</b>) tensile stress and (<b>C</b>) tensile strain.</p>
Full article ">Figure 15
<p>Mass loss of N100-1 and N100-2 at different implantation times in the back of rats.</p>
Full article ">Figure 16
<p>SEM micrographs of the surface of N100-1 (<b>A</b>–<b>C</b>) and N100-2 (<b>D</b>–<b>F</b>) after 0 (<b>A</b>,<b>D</b>); 2 (<b>B</b>,<b>E</b>) and 8 weeks (<b>C</b>,<b>F</b>) implanted in the back of rats. The scale bar was 50 μm.</p>
Full article ">Figure 17
<p>Change in thermal properties of N100-1 and N100-2 networks at different implantation times in the back of rats: (<b>A</b>) <span class="html-italic">T</span>g and (<b>B</b>) <span class="html-italic">T</span>d.</p>
Full article ">Figure 18
<p>The changes in the mechanical properties of N100-1 and N100-2 networks during <span class="html-italic">in vivo</span> degradation: (<b>A</b>) Young’s modulus (<b>B</b>) tensile stress and (<b>C</b>) tensile strain.</p>
Full article ">Figure 19
<p>Histological sections of the tissue surrounding the implant at different stages of <span class="html-italic">in vivo</span> degradation: (<b>A</b>) 2 weeks at 40× magnification; (<b>B</b>) 2 weeks at 100× magnification; (<b>C</b>) 4 weeks at 100× magnification and (<b>D</b>) 12 weeks at 100× magnification.</p>
Full article ">
5857 KiB  
Article
Processing of Polysulfone to Free Flowing Powder by Mechanical Milling and Spray Drying Techniques for Use in Selective Laser Sintering
by Nicolas Mys, Ruben Van De Sande, An Verberckmoes and Ludwig Cardon
Polymers 2016, 8(4), 150; https://doi.org/10.3390/polym8040150 - 19 Apr 2016
Cited by 34 | Viewed by 11016
Abstract
Polysulfone (PSU) has been processed into powder form by ball milling, rotor milling, and spray drying technique in an attempt to produce new materials for Selective Laser Sintering purposes. Both rotor milling and spray drying were adept to make spherical particles that can [...] Read more.
Polysulfone (PSU) has been processed into powder form by ball milling, rotor milling, and spray drying technique in an attempt to produce new materials for Selective Laser Sintering purposes. Both rotor milling and spray drying were adept to make spherical particles that can be used for this aim. Processing PSU pellets by rotor milling in a three-step process resulted in particles of 51.8 μm mean diameter, whereas spray drying could only manage a mean diameter of 26.1 μm. The resulting powders were characterized using Differential Scanning Calorimetry (DSC), Gel Permeation Chromatography (GPC) and X-ray Diffraction measurements (XRD). DSC measurements revealed an influence of all processing techniques on the thermal behavior of the material. Glass transitions remained unaffected by spray drying and rotor milling, yet a clear shift was observed for ball milling, along with a large endothermic peak in the high temperature region. This was ascribed to the imparting of an orientation into the polymer chains due to the processing method and was confirmed by XRD measurements. Of all processed powder samples, the ball milled sample was unable to dissolve for GPC measurements, suggesting degradation by chain scission and subsequent crosslinking. Spray drying and rotor milling did not cause significant degradation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Viscosity curve of PSU solutions in <span class="html-italic">N</span>,<span class="html-italic">N</span>-Dimethylformamide (DMF) as a function of mass percent of PSU measured at 25 °C. The two-fluid nozzle theoretically only allows the spraying of solutions lower than 15 wt % (solid line).</p>
Full article ">Figure 2
<p>(<b>a</b>) Micrograph of spray-dried PSU from a 12 wt % solution in DMF; (<b>b</b>) insets show a magnification of a collapsed structure (above) and string-like structures (below).</p>
Full article ">Figure 3
<p>(<b>a</b>) Microscopic image of PSU taken after 10 min ball milling with (<b>b</b>) enlargement by SEM of the fractionated powder.</p>
Full article ">Figure 4
<p>Schematics of the rotor milling process with the three-step refinement. After milling, the powders are sieved and the redundant powder is re-fed to the rotor miller for further refinement.</p>
Full article ">Figure 5
<p>(<b>a</b>) Coarse powder by pulverization of pellets at 500 μm; (<b>b</b>) First refinement step by pulverisation of coarse powder at 120 μm; (<b>c</b>) Second refinement step by further pulverization of the refined powder until 80 μm and (<b>d</b>) Final powder after additional sieving at 80 μm.</p>
Full article ">Figure 6
<p>Particle Size Distribution (PSD) of the different processed samples. Spray-Dried (<b>SD</b>) sample at best parameter settings, Rotor Milled (<b>RM</b>) powder subjected to the three-step refinement process with final sieving step, and Ball Milled (<b>BM</b>) sample after 10 min (fractionated fine powder).</p>
Full article ">Figure 7
<p>GPC measurements on unprocessed PSU, Rotor Milled powder subjected to the three step refinement process with additional sieving at 80 μm and spray-dried PSU obtained at the best parameter settings. The second measurements are represented in a lighter shade of the representative color.</p>
Full article ">Figure 8
<p>First heating run of DSC measurements on (<b>A</b>) spray-dried powder obtained at best parameter settings; (<b>B</b>) rotor milled powder subjected to the three step refinement process with additional sieving at 80 μm; (<b>C</b>) virgin PSU and (<b>D</b>) fine powder obtained after 10 min of ball milling the PSU pellets.</p>
Full article ">Figure 9
<p>Second heating run of DSC measurements on (<b>A</b>) spray-dried powder obtained at best parameter settings, (<b>B</b>) rotor milled powder subjected to the three step refinement process with additional sieving at 80 μm, (<b>C</b>) virgin PSU and (<b>D</b>) fine powder obtained after 10 min of ball milling the PSU pellets.</p>
Full article ">Figure 10
<p>XRD diffractograms of the fine powder obtained after 10 min of ball milling, rotor milled powder subjected to the three step refinement process with additional sieving at 80 μm, and unprocessed PSU pellets. Small peaks marked with an asterisk are believed to be attributed to the orientation induced in PSU by the milling method.</p>
Full article ">Figure 11
<p>Evaporation model that illustrates possible morphology occurrence depending on parameters used.</p>
Full article ">Figure 12
<p>Possible degradation mechanisms of polysulfone given: (<b>a</b>) crosslinking by phenylation of phenyl radical formed by cleavage of C–S bond; (<b>b</b>) intramolecular phenylation by hydrogen abstraction by the phenyl radical; (<b>c</b>) β-scission of isopropylidene radical following H-abstraction [<a href="#B31-polymers-08-00150" class="html-bibr">31</a>,<a href="#B33-polymers-08-00150" class="html-bibr">33</a>,<a href="#B35-polymers-08-00150" class="html-bibr">35</a>].</p>
Full article ">
2762 KiB  
Article
2H Solid-State NMR Analysis of the Dynamics and Organization of Water in Hydrated Chitosan
by Fenfen Wang, Rongchun Zhang, Tiehong Chen and Pingchuan Sun
Polymers 2016, 8(4), 149; https://doi.org/10.3390/polym8040149 - 19 Apr 2016
Cited by 12 | Viewed by 6944
Abstract
Understanding water–biopolymer interactions, which strongly affect the function and properties of biopolymer-based tissue engineering and drug delivery materials, remains a challenge. Chitosan, which is an important biopolymer for the construction of artificial tissue grafts and for drug delivery, has attracted extensive attention in [...] Read more.
Understanding water–biopolymer interactions, which strongly affect the function and properties of biopolymer-based tissue engineering and drug delivery materials, remains a challenge. Chitosan, which is an important biopolymer for the construction of artificial tissue grafts and for drug delivery, has attracted extensive attention in recent decades, where neutralization with an alkali solution can substantially enhance the final properties of chitosan films cast from an acidic solution. In this work, to elucidate the effect of water on the properties of chitosan films, we investigated the dynamics and different states of water in non-neutralized (CTS-A) and neutralized (CTS-N) hydrated chitosan by mobility selective variable-temperature (VT) 2H solid-state NMR spectroscopy. Four distinct types of water exist in all of the samples with regards to dynamic behavior. First, non-freezable, rigid and strongly bound water was found in the crystalline domain at low temperatures. The second component consists of weakly bound water, which is highly mobile and exhibits isotropic motion, even below 260 K. Another type of water undergoes well-defined 180° flips around their bisector axis. Moreover, free water is also present in the films. For the CTS-A sample in particular, another special water species were bounded to acetic acid molecules via strong hydrogen bonding. In the case of CTS-N, the onset of motions of the weakly bound water molecules at 260 K was revealed by 2H-NMR spectroscopy. This water is not crystalline, even below 260 K, which is also the major contribution to the flexibility of chitosan chains and thus toughness of materials. By contrast, such motion was not observed in CTS-A. On the basis of the 2H solid-state NMR results, it is concluded that the unique toughness of CTS-N mainly originates from the weakly bound water as well as the interactions between water and the chitosan chains. Full article
(This article belongs to the Collection Polysaccharides)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD patterns of CTS-A and CTS-N at room temperature; and (<b>b</b>) stress–strain curves of CTS-A and CTS-N.</p>
Full article ">Figure 2
<p>Dynamic-editing <sup>2</sup>H NMR spectra of CTS-A and CTS-N at room temperature: (<b>a</b>) solid-echo spectra recorded with a repetition time of 2 s; (<b>b</b>) saturation Recovery quadrupolar echo spectra with a saturation recovery delay τ<sub>1</sub> = 10 ms to selectively observe signals of mobile components; and (<b>c</b>) Torchia’s T<sub>1</sub>-filter spectra to selectively observe signals of rigid components.</p>
Full article ">Figure 3
<p>Dynamic-editing VT <sup>2</sup>H NMR spectra for CTS-N obtained from: (<b>a</b>) quadrupolar echo; (<b>b</b>) saturation recovery quadrupolar echo; and (<b>c</b>) Torchia’s T<sub>1</sub> filter quadrupolar echo experiments, respectively.</p>
Full article ">Figure 4
<p>VT <sup>2</sup>H spectra obtained from: (<b>a</b>,<b>c</b>) quadrupolar echo experiments to obtain signals of all components; and (<b>b</b>,<b>d</b>) saturation recovery quadrupolar echo experiments to obtain signals of mobile components for CTS-A (<b>a</b>,<b>b</b>) and CTS-N (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">Figure 5
<p>Line width at half height of the narrow components of the solid-echo spectra for CTS-N (<a href="#polymers-08-00149-f004" class="html-fig">Figure 4</a>).</p>
Full article ">Figure 6
<p>Integral signal intensity of the saturation recovery quadrupolar echo <sup>2</sup>H spectra from <a href="#polymers-08-00149-f004" class="html-fig">Figure 4</a>b,d. The onset of motion of the weakly bound matrix water is shown at 260 K.</p>
Full article ">Figure 7
<p><sup>2</sup>H NMR spectra of unstretched (<b>a</b>) and stretched (<b>b</b>) CTS-N sample. The top spectra correspond to the randomly oriented powder samples, while the bottom spectra correspond to sample placed along the magic angle.</p>
Full article ">
1564 KiB  
Article
Mesogenic Polyelectrolyte Gels Absorb Organic Solvents and Liquid Crystalline Molecules
by Yusuke Nishikori, Kazuya Iseda, Kenta Kokado and Kazuki Sada
Polymers 2016, 8(4), 148; https://doi.org/10.3390/polym8040148 - 19 Apr 2016
Cited by 9 | Viewed by 6149
Abstract
In this paper, mesogenic polyelectrolyte gels (MPEgels) tethering mesogenic groups on the side chains were synthesized from a mesogenic monomer and ionic monomer via a conventional radical polymerization process. The obtained MPEgels absorbed various organic solvents in a wide range of dielectric constants [...] Read more.
In this paper, mesogenic polyelectrolyte gels (MPEgels) tethering mesogenic groups on the side chains were synthesized from a mesogenic monomer and ionic monomer via a conventional radical polymerization process. The obtained MPEgels absorbed various organic solvents in a wide range of dielectric constants from chloroform (ε = 7.6) to DMSO (ε = 46.5). The electrostatic repulsion among the polymer chains and the osmotic pressure between the interior and exterior of the MPEgel is responsible for the high swelling ability, revealed by the common ion effect using tetra(n-hexyl)ammonium tetra(3,5-bis(trifluoromethyl)phenylborate (THATFPB). The obtained MPEgels could also absorb liquid crystalline molecules such as 4-cyano-4’-pentylbiphenyl (5CB), analogously caused by the above-mentioned polyelectrolyte characteristic. The MPEgels exhibited liquid crystal transition temperature (TNI) on differential scanning calorimetry (DSC) measurement, and the increase of the ionic group content lowered TNI. The MPEgels absorbing liquid crystalline molecules exhibited differing TNI, dependent on the compatibility of the mesogenic group on the side chain to the liquid crystalline molecule. Full article
(This article belongs to the Collection Polyelectrolytes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Swelling degrees of (<b>a</b>) <b>MPEG6-<span class="html-italic">p</span></b> and (<b>b</b>) <b>MPEG5-<span class="html-italic">p</span></b> in various organic solvents.</p>
Full article ">Figure 2
<p>Swelling degree of (<b>a</b>) <b>MPEG6-<span class="html-italic">p</span></b> and (<b>b</b>) <b>MPEG5-<span class="html-italic">p</span></b> in a liquid crystalline molecule <b>5CB</b> with and without <b>THATFPB</b>.</p>
Full article ">Figure 3
<p>DSC thermograms of the MPEgels (<b>a</b>) <b>MPEG6-<span class="html-italic">p</span></b> and (<b>b</b>) <b>MPEG5-<span class="html-italic">p</span></b> upon the cooling process (5 °C/min).</p>
Full article ">Scheme 1
<p>Synthetic route for mesogenic monomers (<b>M5</b> and <b>M6</b>).</p>
Full article ">Scheme 2
<p>Preparation of the MPEgels (<b>MPEG5-<span class="html-italic">p</span></b> and <b>MPEG6-<span class="html-italic">p</span></b>). HDA: 1,6-hexanediol diacrylate.</p>
Full article ">
1985 KiB  
Article
Synthesis of Monodisperse Silica Particles Grafted with Concentrated Ionic Liquid-Type Polymer Brushes by Surface-Initiated Atom Transfer Radical Polymerization for Use as a Solid State Polymer Electrolyte
by Takashi Morinaga, Saika Honma, Takeo Ishizuka, Toshio Kamijo, Takaya Sato and Yoshinobu Tsujii
Polymers 2016, 8(4), 146; https://doi.org/10.3390/polym8040146 - 16 Apr 2016
Cited by 17 | Viewed by 9324
Abstract
A polymerizable ionic liquid, N,N-diethyl-N-(2-methacryloylethyl)-N-methylammonium bis(trifluoromethylsulfonyl)imide (DEMM-TFSI), was polymerized via copper-mediated atom transfer radical polymerization (ATRP). The polymerization proceeded in a living manner producing well-defined poly(DEMM-TFSI) of target molecular weight up to about 400 K (including [...] Read more.
A polymerizable ionic liquid, N,N-diethyl-N-(2-methacryloylethyl)-N-methylammonium bis(trifluoromethylsulfonyl)imide (DEMM-TFSI), was polymerized via copper-mediated atom transfer radical polymerization (ATRP). The polymerization proceeded in a living manner producing well-defined poly(DEMM-TFSI) of target molecular weight up to about 400 K (including a polycation and an counter anion). The accurate molecular weight as determined by a GPC analysis combined with a light scattering measurement, and the molecular weight values obtained exhibited good agreement with the theoretical values calculated from the initial molar ratio of DEMM-TFSI and the monomer conversion. Surface-initiated ATRP on the surface of monodisperse silica particles (SiPs) with various diameters was successfully performed, producing SiPs grafted with well-defined poly(DEMM-TFSI) with a graft density as high as 0.15 chains/nm2. Since the composite film made from the silica-particle-decorated polymer brush and ionic liquid shows a relatively high ionic conductivity, we have evaluated the relationship between the grafted brush chain length and the ionic conductivity. Full article
(This article belongs to the Collection Silicon-Containing Polymeric Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of <span class="html-italic">N</span>,<span class="html-italic">N</span>-diethyl<span class="html-italic">-N</span>-(2-methacryloylethyl)-<span class="html-italic">N</span>-methylammonium bis(trifluoromethylsulfonyl)imide (DEMM-TFSI).</p>
Full article ">Figure 2
<p>Plot of ln([<span class="html-italic">M</span>]<sub>0</sub>/[<span class="html-italic">M</span>]) <span class="html-italic">vs. t</span> for the solution polymerization of <span class="html-italic">N</span>,<span class="html-italic">N</span>-diethyl<span class="html-italic">-N</span>-(2-methacryloylethyl)-<span class="html-italic">N</span>-methylammonium bis(trifluoromethylsulfonyl)-imide (DEMM-TFSI, 70 wt %) in acetonitrile at 70 °C: [DEMM-TFSI]<sub>0</sub>/[ethyl 2-bromoisobutyrate]<sub>0</sub>/[Cu(I)Cl]<sub>0</sub>/[Cu(II)Cl<sub>2</sub>]<sub>0</sub>/[2,2′-bipyridine]<sub>0</sub> = 1000/1/4.5/0.5/10.</p>
Full article ">Figure 3
<p>Evolution of number-average molecular weight (<span class="html-italic">M</span><sub>n</sub>) and polydispersity index (<span class="html-italic">M</span><sub>w</sub>/<span class="html-italic">M</span><sub>n</sub>) of poly(DEMM-TFSI) estimated from GPC calibrated by standard poly(ethyleneoxide)s (filled circle) and by MALLS (open circle), as a function of monomer conversion for the solution polymerization of <span class="html-italic">N</span>,<span class="html-italic">N</span>-diethyl<span class="html-italic">-N</span>-(2-methacryloylethyl)-<span class="html-italic">N</span>-methylammonium bis(trifluoromethylsulfonyl)imide (DEMM-TFSI, 70 wt %) in acetonitrile at 70 °C: [DEMM-TFSI]<sub>0</sub>/[ethyl-2-bromoisobutyrate]<sub>0</sub>/[Cu(I)Cl]<sub>0</sub>/[Cu(II)Cl<sub>2</sub>]<sub>0</sub>/[2,2′-bipyridine]<sub>0</sub> = 1000/1/4.5/0.5/10. The broken line represents the theoretical molecular weight (<span class="html-italic">M</span><sub>n(theo)</sub>) values calculated with the initial molar ratio of DEMM-TFSI and the monomer conversion.</p>
Full article ">Figure 4
<p>Calibration curves for the molecular weight calculation estimated with standard poly(ethyleneoxide)s (black circle) and molecular weights of poly(DEMM-TFSI) obtained from GPC-MALLS (red line). Molecular weights of poly(DEMM-TFSI) was calculated by the intercept of Zimm plots from the each SLS measurement at respective elution times (blue, pink, and green dots).</p>
Full article ">Figure 5
<p>Schematic representation of the synthesis of SiP grafted with poly(DEMM-TFSI) brushes by surface-initiated ATRP.</p>
Full article ">Figure 6
<p>Fourier transform-infrared spectra of the products at each step for the fabrication of the PSiPs. The blue and green lines correspond to the hybrid particles after surface-initiated ATRP with number-average molecular weights of the poly(DEMM-TFSI) grafted onto SiPs are 25,600 and 157,000, respectively. The red and black lines corresponds to the pure products of SiP and poly(DEMM-TFSI), respectively. The diameter of the SiP core was 130 nm.</p>
Full article ">Figure 7
<p>Arrhenius plots of ionic conductivity σ (<b>a</b>) for PSiP/IL solid, IL(DEME-TFSI) liquid and bulk poly(DEMM-TFSI) (<span class="html-italic">M</span><sub>n(theo)</sub> = 25,600, 157,000), together with SEM images (<b>b</b>,<b>c</b>) of fractured surfaces of PSiP/IL-solid electrolyte observed at a magnification of 15,000×. In the (<b>a</b>), <b>red</b> open circle, PSiP/IL solid with polymer brush of <span class="html-italic">M</span><sub>n(theo)</sub> = 157,000; <b>red</b> filled circle, PSiP/IL solid with polymer brush of <span class="html-italic">M</span><sub>n(theo)</sub> = 25,600; <b>blue</b> cross, DEME-TFSI (ionic liquid); <b>black</b> filled square, bulk poly(DEMM-TFSI) of <span class="html-italic">M</span><sub>n(theo)</sub> = 25,600; black open square, bulk poly(DEMM-TFSI) of <span class="html-italic">M</span><sub>n(theo)</sub> = 157,000; <b>green</b> filled triangle, bulk poly(DEMM-TFSI) of <span class="html-italic">M</span><sub>n(theo)</sub> = 25,600 including 25% of ionic liquid, DEME-TFSI; <b>green</b> open triangle, bulk poly(DEMM-TFSI) of <span class="html-italic">M</span><sub>n(theo)</sub> = 157,000 including 25% of ionic liquid, DEME-TFSI. SEM image of PSiP/IL solid with polymer brush of <span class="html-italic">M</span><sub>n(theo)</sub> = 25,600 for (<b>b</b>) [<a href="#B19-polymers-08-00146" class="html-bibr">19</a>] and PSiP/IL solid with polymer brush of <span class="html-italic">M</span><sub>n(theo)</sub> = 157,000 for (<b>c</b>).</p>
Full article ">
4109 KiB  
Article
Modification of Spherical Polyelectrolyte Brushes by Layer-by-Layer Self-Assembly as Observed by Small Angle X-ray Scattering
by Yuchuan Tian, Li Li, Haoya Han, Weihua Wang, Yunwei Wang, Zhishuang Ye and Xuhong Guo
Polymers 2016, 8(4), 145; https://doi.org/10.3390/polym8040145 - 15 Apr 2016
Cited by 16 | Viewed by 7034
Abstract
Multilayer modified spherical polyelectrolyte brushes were prepared through alternate deposition of positively charged poly(allylamine hydrochloride) (PAH) and negatively charged poly-l-aspartic acid (PAsp) onto negatively charged spherical poly(acrylic acid) (PAA) brushes (SPBs) on a poly(styrene) core. The charge reversal determined by the [...] Read more.
Multilayer modified spherical polyelectrolyte brushes were prepared through alternate deposition of positively charged poly(allylamine hydrochloride) (PAH) and negatively charged poly-l-aspartic acid (PAsp) onto negatively charged spherical poly(acrylic acid) (PAA) brushes (SPBs) on a poly(styrene) core. The charge reversal determined by the zeta potential indicated the success of layer-by-layer (LBL) deposition. The change of the structure during the construction of multilayer modified SPBs was observed by small-angle X-ray scattering (SAXS). SAXS results indicated that some PAH chains were able to penetrate into the PAA brush for the PAA-PAH double-layer modified SPBs whereas part of the PAH moved towards the outer layer when the PAsp layer was loaded to form a PAA-PAH-PAsp triple-layer system. The multilayer modified SPBs were stable upon changing the pH (5 to 9) and ionic strength (1 to 100 mM). The triple-layer modified SPBs were more tolerated to high pH (even at 11) compared to the double-layer ones. SAXS is proved to be a powerful tool for studying the inner structure of multilayer modified SPBs, which can establish guidelines for the a range of potential applications of multilayer modified SPBs. Full article
(This article belongs to the Collection Polyelectrolytes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the modification of SPBs by layer-by-layer self-assembly.</p>
Full article ">Figure 2
<p>The five-layer model of the distribution of Δρ<sup>e</sup> for multilayer SPB. (<b>a</b>) Schematic illustration of multilayer SPB; (<b>b</b>) Illustration of the five-layer model for multilayer modified SPBs with a core of PS, the red lines, PAH, gray lines, PAA chians.</p>
Full article ">Figure 3
<p>TEM images of (<b>a</b>) PS core; (<b>b</b>) SPB; (<b>c</b>) PAA-PAH double-layer modified SPB; and (<b>d</b>) PAA-PAH-PAsp triple-layer modified SPB.</p>
Full article ">Figure 4
<p>(<b>a</b>) Zeta-potential of multilayer SPBs as a function of the number of deposition layers; (<b>b</b>) DLS curves of multilayer modified SPBs. Symbol denotes: (□, black) SPBs; (○, red) PAA-PAH double-layers; (∆, blue) PAA-PAH-PAsp triple-layers. (pH = 7, C<sub>NaCl</sub> = 10 mM).</p>
Full article ">Figure 5
<p>(<b>a</b>) Scattering intensity of SPBs (○, black), PAA-PAH double-layers (□, red) and PAA-PAH-PAsp triple-layers (∆, blue). The solid line is the fitting curve; (<b>b</b>) The radial profile of the excess electron density of multilayer modified SPBs as a function of the radius. (C<sub>NaCl</sub> = 10 mM, pH = 7). All the SAXS curves were normalized to the same mass fraction of the SPB solution.</p>
Full article ">Figure 6
<p>Effect of pH on the radius of PAA-PAH double-layer modified SPBs measured by DLS and SAXS (C<sub>NaCl</sub> = 10 mM). Symbols denote: (■, black) DLS, (<math display="inline"> <semantics> <mo mathcolor="red">●</mo> </semantics> </math>, red) SAXS.</p>
Full article ">Figure 7
<p>(<b>a</b>) Scattering intensity curves of PAA-PAH double-layer modified SPBs at pH 5 (∆, black), pH 7 (□, blue), and pH 9 (○, red). The solid lines are the corresponding fitting curves; the inset is the enlarged view of scattering curves (0.1 nm<sup>−1</sup> &lt; <span class="html-italic">q</span> &lt; 0.23 nm<sup>−1</sup>); (<b>b</b>) The radial profile of excess electron density of PAA-PAH at different pH (C<sub>NaCl</sub> = 10 mM).</p>
Full article ">Figure 8
<p>Effect of pH on the radius of PAA-PAH-PAsp triple-layer modified SPBs measured by DLS and SAXS (C<sub>NaCl</sub> = 10 mM). Symbols denote: (■, black) DLS, (<math display="inline"> <semantics> <mo mathcolor="red">●</mo> </semantics> </math>, red) SAXS.</p>
Full article ">Figure 9
<p>(<b>a</b>) Scattering intensity curves of PAA-PAH-PAsp triple-layer modified SPBs at pH 5 (∆, black), 7 (□, blue) and 9 (○, red). Solid lines are the fitting curves; (<b>b</b>) The radial profile of excess electron density of PAA-PAH-PAsp triple-layer modified SPBs at different pH (C<sub>NaCl</sub> = 10 mM).</p>
Full article ">Figure 10
<p>Scattering intensity curves of PAA-PAH double-layer modified SPBs at different ionic strengths. Symbols denote: (○, black) 100 mM, (□, red) 10 mM, and (∆, blue) 1 mM. Solid lines are the SAXS fitting data. The inset is the radial profile of excess electron density.</p>
Full article ">Figure 11
<p>(<b>a</b>) Scattering intensity curves of PAA-PAH-PAsp triple-layer modified SPBs at NaCl concentrations of 1 mM (○, black), and 10 mM (□, red). Solid lines represent the fitting curves; (<b>b</b>) Excess electron density of PAA-PAH-PAsp at different salt concentrations.</p>
Full article ">
11107 KiB  
Article
Rapid Hydrophilization of Model Polyurethane/Urea (PURPEG) Polymer Scaffolds Using Oxygen Plasma Treatment
by Rok Zaplotnik, Alenka Vesel, Gregor Primc, Xiangyu Liu, Kevin C. Chen, Chiju Wei, Kaitian Xu and Miran Mozetic
Polymers 2016, 8(4), 144; https://doi.org/10.3390/polym8040144 - 15 Apr 2016
Cited by 3 | Viewed by 7031
Abstract
Polyurethane/urea copolymers based on poly(ethylene glycol) (PURPEG) were exposed to weakly ionized, highly reactive low-pressure oxygen plasma to improve their sorption kinetics. The plasma was sustained with an inductively coupled radiofrequency generator operating at various power levels in either E-mode (up to the [...] Read more.
Polyurethane/urea copolymers based on poly(ethylene glycol) (PURPEG) were exposed to weakly ionized, highly reactive low-pressure oxygen plasma to improve their sorption kinetics. The plasma was sustained with an inductively coupled radiofrequency generator operating at various power levels in either E-mode (up to the forward power of 300 W) or H-mode (above 500 W). The treatments that used H-mode caused nearly instant thermal degradation of the polymer samples. The density of the charged particles in E-mode was on the order of 1016 m−3, which prevented material destruction upon plasma treatment, but the density of neutral O-atoms in the ground state was on the order of 1021 m−3. The evolution of plasma characteristics during sample treatment in E-mode was determined by optical emission spectroscopy; surface modifications were determined by water adsorption kinetics and X-ray photoelectron spectroscopy; and etching intensity was determined by residual gas analysis. The results showed moderate surface functionalization with hydroxyl and carboxyl/ester groups, weak etching at a rate of several nm/s, rather slow activation down to a water contact angle of 30° and an ability to rapidly absorb water. Full article
(This article belongs to the Special Issue Polymers Applied in Tissue Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of Polyurethane/urea copolymers based on poly(ethylene glycol) (PURPEG) polymer.</p>
Full article ">Figure 2
<p>SEM image of an untreated PURPEG sample.</p>
Full article ">Figure 3
<p>Schematic of the experimental setup: 1—plasma reactor, 2—sample, 3—oxygen flask, 4—high pressure valve, 5—leak valve, 6—gate valves, 7—double electrical probe, 8—optical spectrometer, 9—catalytic probe, 10—absolute vacuum gauge, 11—capillary, 12—high vacuum chamber, 13—high vacuum gauge, 14—residual gas analyzer, 15—turbomolecular pump, 16—rotary pumps.</p>
Full article ">Figure 4
<p>An optical spectrum of plasma created in the empty chamber at a forward power of 150 W and a pressure of 33 Pa.</p>
Full article ">Figure 5
<p>OES spectrum of the plasma during the etching of a PURPEG sample with oxygen plasma generated at 900 W forward RF power.</p>
Full article ">Figure 6
<p>OES spectrum after 30 s of plasma treatment at 33 Pa and forward power of 150 W.</p>
Full article ">Figure 7
<p>OES spectrum after 600 s of plasma treatment at 33 Pa and forward power of 150 W.</p>
Full article ">Figure 8
<p>Time evolution of the CO emission peak (519.5 nm), O (777.4 nm) line and Balmer Hα line during sample etching with oxygen plasma generated at a pressure of 33 Pa and 150 W forward RF power.</p>
Full article ">Figure 9
<p>Time evolution of the CO emission peak (519.5 nm) normalized by the O (777.4 nm) line during sample treatment with oxygen plasma; the RF forward power is shown in the key.</p>
Full article ">Figure 10
<p>A mass spectrum measured in an empty chamber filled with oxygen at pressure 33 Pa and forward power of 150 W.</p>
Full article ">Figure 11
<p>A mass spectrum measured in a chamber loaded with PURPEG samples after plasma treatment for 30 s. The pressure was 33 Pa and forward power 150 W.</p>
Full article ">Figure 12
<p>A mass spectrum measured in a chamber loaded with PURPEG samples after plasma treatment for 600 s. The pressure was 33 Pa and forward power 150 W.</p>
Full article ">Figure 13
<p>Time evolution of CO<sub>2</sub>, CO, O<sub>2</sub> and H<sub>2</sub>O partial pressures during samples treatment with oxygen plasma at 33 Pa and 150 W forward RF power.</p>
Full article ">Figure 14
<p>Time evolution of CO during sample treatment with oxygen plasma generated at various forward RF power levels.</p>
Full article ">Figure 15
<p>Time evolution of CO<sub>2</sub>, CO, O<sub>2</sub> and H<sub>2</sub>O partial pressures during sample treatment with oxygen plasma at 33 Pa and 300 W forward RF power.</p>
Full article ">Figure 16
<p>The slope of the oxygen partial pressure during the first few minutes of plasma treatment at 33 Pa and 150 W.</p>
Full article ">Figure 17
<p>The composition of the surface film as calculated from the XPS survey spectra. The pressure was 33 Pa and the forward power was 150 W.</p>
Full article ">Figure 18
<p>Evolution of XPS C1s spectra for PURPEG samples as a result of treatment with oxygen plasma at 33 Pa and 150 W.</p>
Full article ">Figure 19
<p>Water contact angle <span class="html-italic">versus</span> plasma treatment time at 33 Pa and 150 W.</p>
Full article ">Figure 20
<p>Absorption time of a water droplet <span class="html-italic">versus</span> plasma treatment time at 33 Pa and 150 W.</p>
Full article ">
3760 KiB  
Article
Analysis and Testing of Bisphenol A—Free Bio-Based Tannin Epoxy-Acrylic Adhesives
by Shayesteh Jahanshahi, Antonio Pizzi, Ali Abdulkhani and Alireza Shakeri
Polymers 2016, 8(4), 143; https://doi.org/10.3390/polym8040143 - 15 Apr 2016
Cited by 37 | Viewed by 12314
Abstract
A tannin-based epoxy acrylate resin was prepared from glycidyl ether tannin (GET) and acrylic acid. The influence of the reaction condition for producing tannin epoxy acrylate was studied by FT-MIR, 13C-NMR, MALDI-TOF spectroscopy and shear strength. The best reaction conditions for producing [...] Read more.
A tannin-based epoxy acrylate resin was prepared from glycidyl ether tannin (GET) and acrylic acid. The influence of the reaction condition for producing tannin epoxy acrylate was studied by FT-MIR, 13C-NMR, MALDI-TOF spectroscopy and shear strength. The best reaction conditions for producing tannin epoxy acrylate resin without bisphenol A was by reaction between GET and acrylic acid in the presence of a catalyst and hydroquinone at 95 °C for 12 h. FT-MIR, 13C-NMR and MALDI-TOF analysis have confirmed that the resin has been prepared under these conditions. The joints bonded with this resin were tested for block shear strength. The results obtained indicated that the best strength performance was obtained by the bioepoxy-acrylate adhesive resin prepared at 95 °C for a 12-h reaction. Full article
(This article belongs to the Special Issue Renewable Polymeric Adhesives)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The four main structures in commercial flavonoid tannins.</p>
Full article ">Figure 2
<p>FTIR spectrum of the ET3 tannin epoxy acrylic resin sample.</p>
Full article ">Figure 3
<p><sup>13</sup>C-NMR spectrum of the ET3 tannin epoxy acrylate resin.</p>
Full article ">Figure 4
<p>Reaction scheme for the epoxidation of tannin. Formation of epoxy groups on different sites of tannin’s flavonoid units.</p>
Full article ">Figure 5
<p>Reaction scheme of glycidylation of the tannin with acrylic acid (tannin-based epoxy acrylate).</p>
Full article ">Figure 6
<p>MALDI-TOF spectrum of the ET3 sample of tannin epoxy acrylate resin.</p>
Full article ">Scheme 1
<p>Sites numbering of a flavonoid unit.</p>
Full article ">Scheme 2
<p>Epoxy tannin oligomers as previously identified.</p>
Full article ">Scheme 3
<p>Most probable structure of the 476 Da MALDI peak.</p>
Full article ">Scheme 4
<p>Catechin dimer with one epoxy group representing the 654 Da MALDI peak.</p>
Full article ">Scheme 5
<p>Robinetinidin-gallocatechin dimer with one epoxy group equally representing the 654 Da MALDI peak.</p>
Full article ">Scheme 6
<p>Chemical structure describing the 638-639 Da MALDI peak.</p>
Full article ">Scheme 7
<p>Chemical structure describing the 813 Da MALDI peak.</p>
Full article ">Scheme 8
<p>Chemical structure describing the 976 Da MALDI peak.</p>
Full article ">
3960 KiB  
Article
Effect of Rubber Nanoparticle Agglomeration on Properties of Thermoplastic Vulcanizates during Dynamic Vulcanization
by Hanguang Wu, Ming Tian, Liqun Zhang, Hongchi Tian, Youping Wu, Nanying Ning and Guo-Hua Hu
Polymers 2016, 8(4), 127; https://doi.org/10.3390/polym8040127 - 15 Apr 2016
Cited by 37 | Viewed by 8127
Abstract
We previously reported that the dispersed rubber microparticles in ethylene-propylene-diene monomer (EPDM)/polypropylene (PP) thermoplastic vulcanizates (TPVs) are actually agglomerates of rubber nanoparticles. In this study, based on this new understanding of the microstructure of TPV, we further revealed the microstructure-properties relationship of EPDM/PP [...] Read more.
We previously reported that the dispersed rubber microparticles in ethylene-propylene-diene monomer (EPDM)/polypropylene (PP) thermoplastic vulcanizates (TPVs) are actually agglomerates of rubber nanoparticles. In this study, based on this new understanding of the microstructure of TPV, we further revealed the microstructure-properties relationship of EPDM/PP TPV during dynamic vulcanization, especially the effect of the size of rubber nanoparticle agglomerates (dn), the thicknesses of PP ligaments (IDpoly) and the rubber network on the properties of EPDM/PP TPV. We were able to simultaneously obtain a high tensile strength, elongation at break, elastic modulus, and elasticity for the EPDM/PP TPV by the achievement of a smaller dn, a thinner IDpoly and a denser rubber network. Interestingly, the effect of dn and IDpoly on the elastic modulus of EPDM/PP TPV composed of rubber nanoparticle agglomerates is different from that of EPDM/PP TPVs composed of rubber microparticles reported previously. The deformation behavior of the TPVs during stretching was studied to understand the mechanism for the achievement of good mechanical properties. Interestingly, the rubber nanoparticle agglomerates are oriented along the tensile direction during stretching. The TPV samples with smaller and more numerous rubber nanoparticle agglomerates can slow down the development of voids and cracks more effectively, thus leading to increase in tensile strength and elongation at break of the EPDM/PP TPV. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Variations of torque (<span class="html-italic">M</span>), temperature (<span class="html-italic">T</span>), swell ratio (<span class="html-italic">Q</span>) and rubber gel content (% gel) of the ethylene-propylene-diene monomer (EPDM)/polypropylene (PP) (60/40) blend during dynamic vulcanization (DV). Four samples (A, B, C, D) are selected according to the curve.</p>
Full article ">Figure 2
<p>AFM phase images of EPDM/PP (60/40) TPV samples taken at different DV times: (a)–(d) represent the AFM images of samples A–D, respectively. (The lighter and darker regions represent the PP and EPDM phases, respectively).</p>
Full article ">Figure 3
<p>Tension-recovery stress-strain curves of EPDM/PP (60/40) TPV samples taken at different DV times.</p>
Full article ">Figure 4
<p>Hysteresis loss and permanent deformation as a function of interparticle distance (<span class="html-italic">ID<sub>poly</sub></span>).</p>
Full article ">Figure 5
<p>Storage modulus (<span class="html-italic">G’</span>) and complex viscosity (η<span class="html-italic">*</span>) at 210 °C as a function of angular frequency for samples A to D.</p>
Full article ">Figure 6
<p>Tensile properties of EPDM/PP (60/40) TPV samples selected at different DV times: (<b>a</b>) stress-strain curves; (<b>b</b>) elastic modulus.</p>
Full article ">Figure 7
<p>AFM images of elongated samples at different tensile strains (The darker and lighter regions represent the EPDM and PP phases, respectively.): (<b>A</b>) Low magnification; (<b>B</b>) high magnification. a, b, c are the AFM images of Samples A with strain of 200%, 400%, 800%, respectively; d and e are the AFM images of Samples D with strain of 200% and 400%, respectively.</p>
Full article ">Figure 8
<p>SEM micrographs of tensile fractured surfaces of the samples A and D.</p>
Full article ">
1832 KiB  
Article
Density Functional Theory of Polymer Structure and Conformations
by Zhaoyang Wei, Nanying Ning, Liqun Zhang, Ming Tian and Jianguo Mi
Polymers 2016, 8(4), 121; https://doi.org/10.3390/polym8040121 - 15 Apr 2016
Cited by 12 | Viewed by 8606
Abstract
We present a density functional approach to quantitatively evaluate the microscopic conformations of polymer chains with consideration of the effects of chain stiffness, polymer concentration, and short chain molecules. For polystyrene (PS), poly(ethylene oxide) (PEO), and poly(methyl methacrylate) (PMMA) melts with low-polymerization degree, [...] Read more.
We present a density functional approach to quantitatively evaluate the microscopic conformations of polymer chains with consideration of the effects of chain stiffness, polymer concentration, and short chain molecules. For polystyrene (PS), poly(ethylene oxide) (PEO), and poly(methyl methacrylate) (PMMA) melts with low-polymerization degree, as chain length increases, they display different stretching ratios and show non-universal scaling exponents due to their different chain stiffnesses. In good solvent, increase of PS concentration induces the decline of gyration radius. For PS blends containing short (m1 = 1 100) and long (m = 100) chains, the expansion of long chains becomes unobvious once m 1 is larger than 40, which is also different to the scaling properties of ideal chain blends. Full article
(This article belongs to the Special Issue Computational Modeling and Simulation in Polymer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of the polymer model used in this work. Here, a middle segment from a polymer chain (filled black sphere) is fixed at the origin. The density distributions of segments from the tethered fragments (<math display="inline"> <semantics> <mi>C</mi> </semantics> </math> and <math display="inline"> <semantics> <mi>D</mi> </semantics> </math>), free (<math display="inline"> <semantics> <mi>F</mi> </semantics> </math>), and short (<math display="inline"> <semantics> <mi>S</mi> </semantics> </math>) chain molecules are related to the intra- and intermolecular segment–segment correlation functions. In a homogeneous polymer melt, the short chain no longer exists.</p>
Full article ">Figure 2
<p>Comparison of the average (<b>a</b>) inter- and (<b>b</b>) intramolecular correlation functions of the PS chain obtained from theory and molecular simulations [<a href="#B42-polymers-08-00121" class="html-bibr">42</a>] at 413.2 K.</p>
Full article ">Figure 3
<p>(<b>a</b>) Gyration radius as a function of chain length for PS chains at 500 K; (<b>b</b>) Determination of the scaling exponents. The circles are calculated results and the solid line is the best linear regression of the circles.</p>
Full article ">Figure 4
<p>(<b>a</b>) Comparison between experimental data of the form factors and calculated ones for PS melt with the molecular weight 1 × 10<sup>4</sup> in θ solvent; (<b>b</b>) Form factor for different lengths of PS melt at 500 K. Solid lines correspond to DFT calculations in Equation (21). Symbols show the curve obtained from the Debye expression.</p>
Full article ">Figure 5
<p>Structure factor of PS chain with three values of polymerization degree at 500 K.</p>
Full article ">Figure 6
<p>(<b>a</b>) Gyration radius as a function of chain length for PEO and PMMA melts at 300 K; (<b>b</b>) Determination of the scaling exponents for the two polymers. The squares and circles are calculated results and the solid line is the best linear regression.</p>
Full article ">Figure 7
<p>Gyration radius of PS chain as a function of packing fraction at 500 K. The chain length is fixed at <math display="inline"> <semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 8
<p>Gyration radius of long chain PS (<math display="inline"> <semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math>) as a function of chain length of short chain PS (<math display="inline"> <semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>−</mo> <mn>100</mn> </mrow> </semantics> </math>) at 500 K.</p>
Full article ">
3322 KiB  
Article
pH-Triggered Sheddable Shielding System for Polycationic Gene Carriers
by Jialiang Xia, Huayu Tian, Jie Chen, Lin Lin, Zhaopei Guo, Bing Han, Hongyan Yang and Zongcai Feng
Polymers 2016, 8(4), 141; https://doi.org/10.3390/polym8040141 - 14 Apr 2016
Cited by 5 | Viewed by 5715
Abstract
For improving the therapeutic efficiency of tumors and decreasing undesirable side effects, ternary complexes were developed by coating pH-sensitive PEG-b-PLL-g-succinylsulfathiazole (hereafter abbreviated as PPSD) with DNA/PEI polyplexes via electrostatic interaction. PPSD can efficiently shield the surface charge of DNA/PEI. [...] Read more.
For improving the therapeutic efficiency of tumors and decreasing undesirable side effects, ternary complexes were developed by coating pH-sensitive PEG-b-PLL-g-succinylsulfathiazole (hereafter abbreviated as PPSD) with DNA/PEI polyplexes via electrostatic interaction. PPSD can efficiently shield the surface charge of DNA/PEI. The gene transfection efficiency of ternary complexes was lower than that of DNA/PEI at pH 7.4; however, it recovered to the same level as that of DNA/PEI at pH 6.0, attributed to the pH-triggered release of DNA/PEI from ternary complexes. Cell uptake results also exhibited the same trend as transfection at different pH values. The suitable ability for pH-triggered shielding/deshielding estimated that PPSD demonstrates potential as a shielding system for use in in vivo gene delivery. Full article
(This article belongs to the Special Issue Functional Polymers for Medical Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>pH sensitivity of PPSD: (<b>A</b>) pH buffer capacity titration of PPSD; (<b>B</b>) Zeta potential of PPSD solution at different pH.</p>
Full article ">Figure 2
<p>Surface charge (<b>A</b>) and particle size; (<b>B</b>) characterization of DNA/PEI/PPSD ternary complexes at pH 7.4. Lane 1 to 4, DNA/PEI/PPSD at ratios (<span class="html-italic">wt/wt/wt</span>) of 1:1:0, 1:1:8, 1:1:16, and 1:1:32, respectively.</p>
Full article ">Figure 3
<p>Particle size (<b>A</b>) and zeta potential (<b>B</b>) of DNA/PEI/PPSD (1:1:32) ternary complexes at various pHs. The solid bars could be the DNA/PEI complex and the open bars may be the aggregates of neutral PPSD in the (<b>A</b>).</p>
Full article ">Figure 4
<p>Gel retardation assay. Lane 1: DNA; lane 2 to 5, DNA/PEI/PPSD at ratios (<span class="html-italic">wt/wt/wt</span>) of 1:1:0, 1:1:8, 1:1:16, 1:1:32, respectively.</p>
Full article ">Figure 5
<p>Cytotoxicity of PPSD.</p>
Full article ">Figure 6
<p>Transfection efficiency of the ternary complex at pH 7.4 and 6.0 in HeLa cells. *, <span class="html-italic">p</span> &gt; 0.05; **, 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05. ***, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>CLSM images of HeLa cells incubated with DNA/PEI (<span class="html-italic">wt</span>/<span class="html-italic">wt</span> = 1:1) or DNA/PEI/PPSD (<span class="html-italic">wt</span>/<span class="html-italic">wt</span>/<span class="html-italic">wt</span> = 1:1:32) at pH 7.4 or 6.0. DAPI, cell nucleus (blue); CY5-pDNA (red). Scale bars = 20 μm.</p>
Full article ">Figure 8
<p>Cellular uptake of DNA/PEI (1/1) (<b>A</b>) and DNA/PEI/PPSD (1/1/32); (<b>B</b>) at different pH values in HeLa cells (gray, control; red, pH 7.4; green, pH 6.8) (All complexing ratios are expressed as weight ratios).</p>
Full article ">Scheme 1
<p>Synthesis of PPSD (<b>A</b>) and the pH-sensitive mechanism of sulfonamide units (<b>B</b>).</p>
Full article ">
3043 KiB  
Article
Transparent Blend of Poly(Methylmethacrylate)/Cellulose Acetate Butyrate for the Protection from Ultraviolet
by Raouf Mahmood Raouf, Zaidan Abdul Wahab, Nor Azowa Ibrahim, Zainal Abidin Talib and Buong Woei Chieng
Polymers 2016, 8(4), 128; https://doi.org/10.3390/polym8040128 - 14 Apr 2016
Cited by 17 | Viewed by 7646
Abstract
The use of transparent polymers as an alternative to glass has become widespread. However, the direct exposure of these materials to climatic conditions of sunlight and heat decrease the lifetime cost of these products. The aim of this study was to minimize the [...] Read more.
The use of transparent polymers as an alternative to glass has become widespread. However, the direct exposure of these materials to climatic conditions of sunlight and heat decrease the lifetime cost of these products. The aim of this study was to minimize the harm caused by ultraviolet (UV) radiation exposure to transparent poly(methylmethacrylate) (PMMA), which usually leads to changes in the physical and chemical properties of these materials and reduced performance. This was achieved using environmentally friendly cellulose acetate butyrate (CAB). The optical, morphological, and thermal properties of CAB blended with transparent PMMA was studied using UV-VIS spectrophotometry, scanning electron microscopy, X-ray diffraction, dynamic mechanical analysis, and thermal gravimetric analysis. The results show that CAB was able to reduce the effects of UV radiation by making PMMA more transparent to UV light, thereby preventing the negative effects of trapped radiation within the compositional structure, while maintaining the amorphous structure of the blend. The results also show that CAB blended with PMMA led to some properties commensurate with the requirements of research in terms of a slight increase in the value of the modulus and the glass transition temperature for the PMMA/CAB blend. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Absorbance of PMMA and CAB in different concentrations at 226 nm; (<b>b</b>) transmittance of PMMA and CAB in different concentrations at 798 nm; (<b>c</b>) PMMA/CAB absorbance and transmittance spectrum with concentration; and (<b>d</b>) PMMA/CAB sheets in different concentrations.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Absorbance of PMMA and CAB in different concentrations at 226 nm; (<b>b</b>) transmittance of PMMA and CAB in different concentrations at 798 nm; (<b>c</b>) PMMA/CAB absorbance and transmittance spectrum with concentration; and (<b>d</b>) PMMA/CAB sheets in different concentrations.</p>
Full article ">Figure 2
<p>SEM images for PMMA and PMMA/10%CAB blend with different scale bars: (<b>a</b>) PMMA/10 µm; (<b>b</b>) PMMA/5 µm; (<b>c</b>) PMMA/2 µm; (<b>d</b>) PMMA/10%CAB/10 µm; (<b>e</b>) PMMA/10%CAB/5 µm and (<b>f</b>) PMMA/10%CAB/2 µm.</p>
Full article ">Figure 3
<p>XRD pattern for: (<b>a</b>) pure PMMA, and (<b>b</b>) PMMA/10%CAB.</p>
Full article ">Figure 4
<p>Pure PMMA and PMMA/10%CAB trace for (<b>a</b>) storage modulus E'; (<b>b</b>) loss modulus E"; and (<b>c</b>) the loss factor (tan δ).</p>
Full article ">Figure 5
<p>TGA- trace for (<b>a</b>) pure PMMA; and (<b>b</b>) PMMA/10%CAB.</p>
Full article ">
1166 KiB  
Article
Enzyme-Catalyzed Synthesis of Water-Soluble Conjugated Poly[2-(3-thienyl)-Ethoxy-4-Butylsulfonate]
by Yun Zhao, Hongyan Zhu, Xinyang Wang, Yingying Liu, Xiang Wu, Heyuan Zhou and Zhonghai Ni
Polymers 2016, 8(4), 139; https://doi.org/10.3390/polym8040139 - 13 Apr 2016
Cited by 6 | Viewed by 6673
Abstract
An environmentally friendly water-soluble conjugated polythiophene poly[2-(3-thienyl)-ethoxy-4-butylsulfonate] (PTEBS) has been found to be effective for making hybrid solar cells. In this work, we first report the enzyme-catalyzed polymerization of (3-thienyl)-ethoxy-4-butylsulfonate (TEBS) using horseradish peroxidase (HRP) enzyme as a catalyst and hydrogen peroxide (H [...] Read more.
An environmentally friendly water-soluble conjugated polythiophene poly[2-(3-thienyl)-ethoxy-4-butylsulfonate] (PTEBS) has been found to be effective for making hybrid solar cells. In this work, we first report the enzyme-catalyzed polymerization of (3-thienyl)-ethoxy-4-butylsulfonate (TEBS) using horseradish peroxidase (HRP) enzyme as a catalyst and hydrogen peroxide (H2O2) as an oxidant in an aqueous buffer. This enzyme-catalyzed polymerization is a “green synthesis process” for the synthesis of water-soluble conjugated PTEBS, the benefits of which include a simple setting, high yields, and an environmentally friendly route. Fourier transform infrared spectra (FTIR) and UV–Vis absorption spectra confirm the successful enzyme-catalyzed polymerization of TEBS. The thermo gravimetric (TG) data show the obtained PTEBS is stable over a fairly high range of temperatures. The present PTEBS has a good solubility in water and ethanol, and photoluminescence quenching of PTEBS/titanium dioxide (TiO2) composite implies that the excitons dissociate and separate successfully at the interface of PTEBS and TiO2, which help to build solar cells using green processing methods. Full article
(This article belongs to the Special Issue Enzymatic Polymer Synthesis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of enzyme-catalyzed polymerization of water-soluble (3-thienyl)-ethoxy-4-butylsulfonate (TEBS) with horseradish peroxidase (HRP).</p>
Full article ">Figure 2
<p>Fourier transform infrared (FTIR) spectra of monomer TEBS and polymer poly[2-(3-thienyl)-ethoxy-4-butylsulfonate] (PTEBS).</p>
Full article ">Figure 3
<p>UV–Vis spectra of monomer TEBS and polymer PTEBS.</p>
Full article ">Figure 4
<p>TG curve of polymer PTEBS heated from 0 to 600 °C.</p>
Full article ">Figure 5
<p>Photoluminescence of PTEBS/TiO<sub>2</sub> water solution.</p>
Full article ">
3983 KiB  
Review
Supported Catalysts Useful in Ring-Closing Metathesis, Cross Metathesis, and Ring-Opening Metathesis Polymerization
by Jakkrit Suriboot, Hassan S. Bazzi and David E. Bergbreiter
Polymers 2016, 8(4), 140; https://doi.org/10.3390/polym8040140 - 12 Apr 2016
Cited by 31 | Viewed by 9548
Abstract
Ruthenium and molybdenum catalysts are widely used in synthesis of both small molecules and macromolecules. While major developments have led to new increasingly active catalysts that have high functional group compatibility and stereoselectivity, catalyst/product separation, catalyst recycling, and/or catalyst residue/product separation remain an [...] Read more.
Ruthenium and molybdenum catalysts are widely used in synthesis of both small molecules and macromolecules. While major developments have led to new increasingly active catalysts that have high functional group compatibility and stereoselectivity, catalyst/product separation, catalyst recycling, and/or catalyst residue/product separation remain an issue in some applications of these catalysts. This review highlights some of the history of efforts to address these problems, first discussing the problem in the context of reactions like ring-closing metathesis and cross metathesis catalysis used in the synthesis of low molecular weight compounds. It then discusses in more detail progress in dealing with these issues in ring opening metathesis polymerization chemistry. Such approaches depend on a biphasic solid/liquid or liquid separation and can use either always biphasic or sometimes biphasic systems and approaches to this problem using insoluble inorganic supports, insoluble crosslinked polymeric organic supports, soluble polymeric supports, ionic liquids and fluorous phases are discussed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of olefin metathesis catalysts <b>1</b>–<b>5</b>.</p>
Full article ">Figure 2
<p>Examples of DVB-PS-supported ( <span class="html-fig-inline" id="polymers-08-00140-i001"> <img alt="Polymers 08 00140 i001" src="/polymers/polymers-08-00140/article_deploy/html/images/polymers-08-00140-i001.png"/></span> = divinylbenzene cross-linked polystyrene or <span class="html-fig-inline" id="polymers-08-00140-i002"> <img alt="Polymers 08 00140 i002" src="/polymers/polymers-08-00140/article_deploy/html/images/polymers-08-00140-i002.png"/></span> = divinylbenzene cross-linked 4-vinylpyridine) olefin metathesis catalysts <b>6–11</b>.</p>
Full article ">Figure 3
<p>Examples of silica-supported olefin metathesis catalysts <b>12</b>–<b>17</b>.</p>
Full article ">Figure 4
<p>Examples of silica-supported olefin metathesis catalysts <b>18</b>–<b>21</b>.</p>
Full article ">Figure 5
<p>Polystyrene-based-supported Schrock catalysts <b>22</b>–<b>26</b> and polynorbornene-based-supported Schrock catalysts <b>27</b>–<b>30</b>.</p>
Full article ">Figure 6
<p>Ionic liquid-tagged ruthenium catalysts <b>33</b>–<b>36</b>.</p>
Full article ">Figure 7
<p>Fluorous-tagged ruthenium catalysts <b>37</b>–<b>41</b>.</p>
Full article ">Figure 8
<p>PEG-supported Hoveyda-Grubbs catalyst <b>42</b> and <b>43</b>.</p>
Full article ">Figure 9
<p>Schematic representation of (<b>a</b>) thermomorphic liquid/liquid separation where a biphasic mixture of <b>polar</b> and <b>nonpolar</b> solvent is heated to form a <b>miscible solvent mixture</b>; (<b>b</b>) a latent biphasic liquid/liquid separation where a room temperature mixture of <b>miscible solvents</b> is used to effect a reaction and the solvent mixture is perturbed (e.g., by addition of another solvent (e.g., water) or some salt) to form a separable biphasic mixture of <b>polar</b> and <b>nonpolar</b> solvents; and (<b>c</b>) a thermomorphic solid/liquid separation system where a mixture of substrate and <b>insoluble polymer-bound catalyst</b> is heated to form a <b>monophasic solution</b> that is in turn cooled to form a <b>precipitate of the polymer bound catalyst</b> that is then separated from the <b>product solution</b> by filtration.</p>
Full article ">Figure 10
<p>Polyolefin-supported Hoveyda-Grubbs catalysts <b>44</b>–<b>46</b>.</p>
Full article ">Figure 11
<p>Polyisobutylene (PIB<sub>1000</sub>)-supported Grubbs second generation catalysts <b>63</b> and Grubbs catalyst <b>64</b>.</p>
Full article ">Figure 12
<p>ROMP products <b>74</b>–<b>78</b>.</p>
Full article ">Scheme 1
<p>Asymmetric Ring-Opening Cross Metathesis Reaction of <b>31</b> catalyzed by <b>22</b>.</p>
Full article ">Scheme 2
<p>DVB-PS-Supported Olefin Metathesis Catalysts <b>6</b>, <b>47</b>, and <b>48</b>.</p>
Full article ">Scheme 3
<p>Synthesis of Monolith-Supported Ruthenium Olefin Metathesis Catalyst <b>50</b>.</p>
Full article ">Scheme 4
<p>Synthesis of Silica-Supported Schrock Catalyst <b>53</b>.</p>
Full article ">Scheme 5
<p>ROMP Reactions of <b>55</b> and <b>57</b> Catalyzed by <b>54</b>.</p>
Full article ">Scheme 6
<p>ROMP Reaction of <b>57</b> Catalyzed by PEG-Supported Hoveyda-Grubbs Catalyst <b>59</b>.</p>
Full article ">Scheme 7
<p>ROMP Reaction of <b>61</b> Catalyzed by PEG-Supported Hoveyda-Grubbs Catalyst <b>60</b>.</p>
Full article ">Scheme 8
<p>ROMP Reactions of <b>65</b>–<b>69</b> Catalyzed by <b>63</b>.</p>
Full article ">
3287 KiB  
Article
A Hyper-Viscoelastic Constitutive Model for Polyurea under Uniaxial Compressive Loading
by Yang Bai, Chunmei Liu, Guangyan Huang, Wei Li and Shunshan Feng
Polymers 2016, 8(4), 133; https://doi.org/10.3390/polym8040133 - 12 Apr 2016
Cited by 44 | Viewed by 8295
Abstract
A hyper-viscoelastic constitutive model for polyurea by separating hyperelastic and viscoelastic behaviors has been put forward. Hyperelasticity represents the rate-independent responses at low strain rates, described by a three-parameter Mooney-Rivlin model and a third Ogden model. By fitting the quasi-static experimental data, the [...] Read more.
A hyper-viscoelastic constitutive model for polyurea by separating hyperelastic and viscoelastic behaviors has been put forward. Hyperelasticity represents the rate-independent responses at low strain rates, described by a three-parameter Mooney-Rivlin model and a third Ogden model. By fitting the quasi-static experimental data, the Ogden model is more appropriate to describe the hyperelastic behaviors for its better agreement at strain over 0.3. Meanwhile, viscoelasticity represents the rate-dependent responses at high strain rates, described by the Standard Linear Solids (SLS) model and the K-BKZ model. By fitting the experimental data of split Hopkinson pressure bar (SHPB), the SLS model is more appropriate to describe the viscoelastic behaviors at strain rates below 1600 s−1, but the K-BKZ model performs better at strain rates over 2100 s−1 because of the substantial increase of Young’s modulus and the state of polyurea transforming from rubbery to glassy. The K-BKZ model is chosen to describe the viscoelastic behavior, for its low Root Mean Square Error (RMSE) at strain rates below 1600 s−1. From the discussion above, the hyper-viscoelastic constitutive model is chosen to be the combination of the Ogden model and the K-BKZ model. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The microstructure of polyurea under SEM.</p>
Full article ">Figure 2
<p>The process propagation of stress waves in the specimen, (<b>a</b>) is the <span class="html-italic">X-t</span> diagram, showing the process of stress wave propagation between bars and the specimen; and (<b>b</b>) is the σ<span class="html-italic">-v</span> diagram, showing the value of stress in the specimen changing with k.</p>
Full article ">Figure 3
<p>The size of the specimen corresponding to the dimensionless stress difference.</p>
Full article ">Figure 4
<p>The high-speed photos of deformation of specimen: (<b>a</b>) before, (<b>b</b>) during, and (<b>c</b>) after.</p>
Full article ">Figure 5
<p>Parallel mechanical response of part A and B.</p>
Full article ">Figure 6
<p>Schematic of the standard linear solid (SLS) model.</p>
Full article ">Figure 7
<p>The true stress-strain curves of polyurea at different strain rates.</p>
Full article ">Figure 8
<p>Young’s modulus of polyurea at different logarithmic strain rates.</p>
Full article ">Figure 9
<p>The failure forms of polyurea at different strain rates: (<b>a</b>) strain rate 900 s<sup>−1</sup>; (<b>b</b>) strain rate 1600 s<sup>−1</sup>; (<b>c</b>) strain rate 2100 s<sup>−1</sup>; (<b>d</b>) strain rate 3000 s<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Comparison between experimental data and theoretical model: (<b>a</b>) quasi-static compression experimental data and Mooney-Rivlin model; (<b>b</b>) quasi-static compression experimental data and Ogden model; (<b>c</b>) SHPB experimental data and SLS model; and (<b>d</b>) SHPB experimental data and K-BKZ model.</p>
Full article ">
5687 KiB  
Article
Preparation of Esterified Bacterial Cellulose for Improved Mechanical Properties and the Microstructure of Isotactic Polypropylene/Bacterial Cellulose Composites
by Bo Wang, Dan Yang, Hai-rong Zhang, Chao Huang, Lian Xiong, Jun Luo and Xin-de Chen
Polymers 2016, 8(4), 129; https://doi.org/10.3390/polym8040129 - 12 Apr 2016
Cited by 27 | Viewed by 6899
Abstract
Bacterial cellulose (BC) has great potential to be used as a new filler to reinforce isotactic polypropylene (iPP) due to its high crystallinity, biodegradability, and efficient mechanical properties. In this study, esterification was used to modify BC, which improved the surface compatibility of [...] Read more.
Bacterial cellulose (BC) has great potential to be used as a new filler to reinforce isotactic polypropylene (iPP) due to its high crystallinity, biodegradability, and efficient mechanical properties. In this study, esterification was used to modify BC, which improved the surface compatibility of the iPP and BC. The results indicated that the cellulose octoate (CO) changed the surface properties from hydrophilic to lipophilic. Compared to the pure iPP, the tensile strength, charpy notched impact strength, and tensile modulus of the iPP/BC composites increased by 9.9%, 7.77%, and 15.64%, respectively. However, the addition of CO reinforced the iPP/CO composites. The tensile strength, charpy notched impact strength, and tensile modulus of the iPP/CO composites increased by 14.23%, 14.08%, and 17.82% compared to the pure iPP. However, the elongation at break of both the composites is decreased. The SEM photographs and particle size distribution of the composites showed improvements when the change of polarity of the BC surface, interface compatibility, and dispersion of iPP improved. Full article
(This article belongs to the Collection Polysaccharides)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The micro topography of bacterial cellulose.</p>
Full article ">Figure 2
<p>Infiltration experiment of BC and cellulose octoate (CO).</p>
Full article ">Figure 3
<p>Fourier transform infrared (FT-IR) spectra of BC and CO.</p>
Full article ">Figure 4
<p>X-ray spectra of BC and CO.</p>
Full article ">Figure 5
<p>Tensile strength of the composites.</p>
Full article ">Figure 6
<p>Tensile modulus of the composites.</p>
Full article ">Figure 7
<p>Elongation at break of the composites.</p>
Full article ">Figure 8
<p>Charpy notched impact strength of the composites.</p>
Full article ">Figure 9
<p>SEM photographs of the composites tensile fractured surface.</p>
Full article ">Figure 10
<p>Particle size distribution of BC or CO in the isotactic polypropylene (iPP) matrix.</p>
Full article ">
4910 KiB  
Review
Precision Synthesis of Functional Polysaccharide Materials by Phosphorylase-Catalyzed Enzymatic Reactions
by Jun-ichi Kadokawa
Polymers 2016, 8(4), 138; https://doi.org/10.3390/polym8040138 - 11 Apr 2016
Cited by 42 | Viewed by 10385
Abstract
In this review article, the precise synthesis of functional polysaccharide materials using phosphorylase-catalyzed enzymatic reactions is presented. This particular enzymatic approach has been identified as a powerful tool in preparing well-defined polysaccharide materials. Phosphorylase is an enzyme that has been employed in the [...] Read more.
In this review article, the precise synthesis of functional polysaccharide materials using phosphorylase-catalyzed enzymatic reactions is presented. This particular enzymatic approach has been identified as a powerful tool in preparing well-defined polysaccharide materials. Phosphorylase is an enzyme that has been employed in the synthesis of pure amylose with a precisely controlled structure. Similarly, using a phosphorylase-catalyzed enzymatic polymerization, the chemoenzymatic synthesis of amylose-grafted heteropolysaccharides containing different main-chain polysaccharide structures (e.g., chitin/chitosan, cellulose, alginate, xanthan gum, and carboxymethyl cellulose) was achieved. Amylose-based block, star, and branched polymeric materials have also been prepared using this enzymatic polymerization. Since phosphorylase shows a loose specificity for the recognition of substrates, different sugar residues have been introduced to the non-reducing ends of maltooligosaccharides by phosphorylase-catalyzed glycosylations using analog substrates such as α-d-glucuronic acid and α-d-glucosamine 1-phosphates. By means of such reactions, an amphoteric glycogen and its corresponding hydrogel were successfully prepared. Thermostable phosphorylase was able to tolerate a greater variance in the substrate structures with respect to recognition than potato phosphorylase, and as a result, the enzymatic polymerization of α-d-glucosamine 1-phosphate to produce a chitosan stereoisomer was carried out using this enzyme catalyst, which was then subsequently converted to the chitin stereoisomer by N-acetylation. Amylose supramolecular inclusion complexes with polymeric guests were obtained when the phosphorylase-catalyzed enzymatic polymerization was conducted in the presence of the guest polymers. Since the structure of this polymeric system is similar to the way that a plant vine twines around a rod, this polymerization system has been named “vine-twining polymerization”. Through this approach, amylose supramolecular network materials were fabricated using designed graft copolymers. Furthermore, supramolecular inclusion polymers were formed by vine-twining polymerization using primer–guest conjugates. Full article
(This article belongs to the Special Issue Enzymatic Polymer Synthesis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Formation of glycosidic linkage between anomeric hydroxy group of sugar residue and hydroxy group of another sugar residue.</p>
Full article ">Figure 2
<p>Reaction scheme for enzymatic glycosylation.</p>
Full article ">Figure 3
<p>Phosphorylase-catalyzed (<b>a</b>) phosphorolysis and glucosylation and (<b>b</b>) enzymatic polymerization.</p>
Full article ">Figure 4
<p>Phosphorylase-catalyzed enzymatic polymerization using modified primer.</p>
Full article ">Figure 5
<p>Phosphorylase-catalyzed enzymatic glycosylations using analogue substrates as glycosyl donors to produce non-natural pentasaccharides.</p>
Full article ">Figure 6
<p>Chemoenzymatic synthesis of amylose-grafted heteropolysaccharides via reductive amination and condensation, followed by phosphorylase-catalyzed enzymatic polymerization.</p>
Full article ">Figure 7
<p>Phosphorylase-catalyzed enzymatic polymerization using glycogen as multifunctional primer to produce hydrogel.</p>
Full article ">Figure 8
<p>Thermostable phosphorylase-catalyzed enzymatic glucuronylation using GlcA-1-P as glycosyl donor to produce acidic tetrasaccharide.</p>
Full article ">Figure 9
<p>Phosphorylase-catalyzed successive enzymatic reactions to produce amphoteric glycogen hydrogel.</p>
Full article ">Figure 10
<p>Differences in glycosylation using Man-1-P/GlcN-1-P as glycosyl donors by potato and thermostable phosphorylase catalyses.</p>
Full article ">Figure 11
<p>Thermostable phosphorylase-catalyzed enzymatic polymerization of α-<span class="html-small-caps">d</span>-glucosamine 1-phosphate to produce chitosan stereoisomer and subsequent <span class="html-italic">N</span>-acetylation to produce chitin stereoisomer.</p>
Full article ">Figure 12
<p>Image of vine-twining polymerization to produce amylose–polymer inclusion complexes and typical guest polymers that have been employed in this system.</p>
Full article ">Figure 13
<p>Parallel enzymatic polymerization system to produce inclusion complex from amylose and strongly hydrophobic polyester.</p>
Full article ">Figure 14
<p>(<b>a</b>) Preparation of amylose supramolecular network materials by vine-twining polymerization using graft copolymers having hydrophilic main-chains and hydrophobic guest graft chains and (<b>b</b>) photographs before and after vine-twining polymerization.</p>
Full article ">Figure 15
<p>Preparation of (<b>a</b>) linear and (<b>b</b>) hyperbranched supramolecular polymers by vine-twining polymerization using primer–guest conjugates.</p>
Full article ">
3905 KiB  
Article
Hg(II) Coordination Polymers Based on N,N’-bis(pyridine-4-yl)formamidine
by Wayne Hsu, Xiang-Kai Yang, Pradhumna Mahat Chhetri and Jhy-Der Chen
Polymers 2016, 8(4), 137; https://doi.org/10.3390/polym8040137 - 11 Apr 2016
Cited by 11 | Viewed by 5989
Abstract
Reactions of N,N’-bis(pyridine-4-yl)formamidine (4-Hpyf) with HgX2 (X = Cl, Br, and I) afforded the formamidinate complex {[Hg(4-pyf)2]·(THF)}n, 1, and the formamidine complexes {[HgX2(4-Hpyf)]·(MeCN)}n (X = Br, 2; I, 3), which have been structurally characterized [...] Read more.
Reactions of N,N’-bis(pyridine-4-yl)formamidine (4-Hpyf) with HgX2 (X = Cl, Br, and I) afforded the formamidinate complex {[Hg(4-pyf)2]·(THF)}n, 1, and the formamidine complexes {[HgX2(4-Hpyf)]·(MeCN)}n (X = Br, 2; I, 3), which have been structurally characterized by X-ray crystallography. Complex 1 is a 2D layer with the {44·62}-sql topology and complexes 2 and 3 are helical chains. While the helical chains of 2 are linked through N–H···Br hydrogen bonds, those of 3 are linked through self-complementary double N–H···N hydrogen bonds, resulting in 2D supramolecular structures. The 4-pyf- ligands of 1 coordinate to the Hg(II) ions through one pyridyl and one adjacent amine nitrogen atoms and the 4-Hpyf ligands of 2 and 3 coordinate to the Hg(II) ions through two pyridyl nitrogen atoms, resulting in new bidentate binding modes. Complexes 1–3 provide a unique opportunity to envisage the effect of the halide anions of the starting Hg(II) salts on folding and unfolding the Hg(II) coordination polymers. Density function theory (DFT) calculation indicates that the emission of 1 is due to intraligand π→π * charge transfer between two different 4-pyf- ligands, whereas those of 2 and 3 can be ascribed to the charge transfer from non-bonding p-type orbitals of the halide anions to π * orbitals of the 4-pyf- ligands (n→π *). The gas sorption properties of the desolvated product of 1 are compared with the Cu analogues to show that the nature of the counteranion and the solvent-accessible volume are important in determining their adsorption capability. Full article
(This article belongs to the Special Issue Coordination Polymers: New Materials for Multiple Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Binding modes of the anionic 4-pyf<sup>−</sup> and neutral 4-Hpyf ligand. (<b>a</b>) Neutral ligand; (<b>b</b>) Bidentate bonding mode through two inner amine nitrogen atoms; (<b>c</b>) Tetradentate bonding mode; (<b>d</b>) Tridentate bonding mode; (<b>e</b>) Bidentate bonding mode through one amine and one pyridyl nitrogen atoms and (<b>f</b>) Bidentate bonding mode through two pyridyl nitrogen atoms.</p>
Full article ">Figure 2
<p>A packing diagram exhibiting the supramolecular structure of 4-Hpyf, showing that the molecules are linked through N–H···N hydrogen bonds.</p>
Full article ">Figure 3
<p>(<b>a</b>) Coordination environment about the Hg(II) ion of <b>1</b>; (<b>b</b>) A 2D layer structure for <b>1</b> showing the outward dangling nitrogen atoms. Thermal ellipsoids are shown at 30% probability level; (<b>c</b>) A packing diagram of <b>1</b> showing the intercalation of THF.</p>
Full article ">Figure 4
<p>(<b>a</b>) A representative coordination environment about the Hg(II) ion for <b>2</b> (X = Br) and <b>3</b> (X = I); Thermal ellipsoids are shown at 30% probability level; (<b>b</b>) A representative drawing showing the 1D helical chain for <b>2</b> and <b>3</b>; (<b>c</b>) A packing diagram for <b>2</b>; (<b>d</b>) A packing diagram for <b>3</b>.</p>
Full article ">Figure 5
<p>The emission and excitation spectra for complexes <b>1</b>–<b>3</b>.</p>
Full article ">Figure 6
<p>Drawings showing the electron contribution of (<b>a</b>) the highest occupied molecular orbital (HOMO) of <b>1</b>, (<b>b</b>) the lowest unoccupied molecular orbital (LUMO) of <b>1</b>, (<b>c</b>) HOMO (bottom) and LUMO (top) of <b>2</b>, and (<b>d</b>) HOMO (bottom) and LUMO (top) of <b>3</b>.</p>
Full article ">Figure 7
<p>The N<sub>2</sub> adsorption isotherms for <b>1’</b>, <b>4’</b>, <b>5a’</b>, and <b>5b’</b>.</p>
Full article ">Figure 8
<p>The H<sub>2</sub> adsorption isotherms for <b>1’</b>, <b>4’</b>, <b>5a’</b>, and <b>5b’</b>.</p>
Full article ">Figure 9
<p>The CO<sub>2</sub> adsorption isotherms for <b>1’</b>, <b>4’</b>, <b>5a’</b>, and <b>5b’</b> at (<b>a</b>) 273 and (<b>b</b>) 298 K.</p>
Full article ">Figure 10
<p>Isosteric heat of adsorption of <b>1’</b>, <b>4’</b>, <b>5a’</b>, and <b>5b’</b> at different CO<sub>2</sub> uptakes.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop