[go: up one dir, main page]

Next Issue
Volume 5, December
Previous Issue
Volume 5, June
 
 
polymers-logo

Journal Browser

Journal Browser

Polymers, Volume 5, Issue 3 (September 2013) – 13 articles , Pages 873-1168

  • 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:
270 KiB  
Article
Enzymatic Synthesis and Characterization of Thermosensitive Polyester with Pendent Ketoprofen
by Wan-Xia Wu, Hai-Yang Wang, Na Wang, Wei-Wei Zhang, Han Lai and Xiao-Qi Yu
Polymers 2013, 5(3), 1158-1168; https://doi.org/10.3390/polym5031158 - 23 Sep 2013
Cited by 6 | Viewed by 6956
Abstract
Three linear polyesters with pendant ketoprofen were synthesized by copolymerization of polyethylene glycol (PEG) with malic acid (thiomalic acid or aspartic acid) using lipase B acrylic resin from Candida antarctica (CAL-B) at 90 °C respectively. These thermosensitive polyesters exhibit a lower critical solution [...] Read more.
Three linear polyesters with pendant ketoprofen were synthesized by copolymerization of polyethylene glycol (PEG) with malic acid (thiomalic acid or aspartic acid) using lipase B acrylic resin from Candida antarctica (CAL-B) at 90 °C respectively. These thermosensitive polyesters exhibit a lower critical solution temperature (LCST) at 10–12 °C. The in vitro study demonstrated that these polyesters could release ketoprofen in neutral and alkaline medium but showed hydrolytic stability in acid medium. These results suggest that, with pendant drugs, these thermosensitive polyesters have potential applications in biomedical materials. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Plot of transmittance as a function of temperature measured for polyester aqueous solution (0.5 wt %) at a heating/cooling rate of 0.5 °C∙min<sup>−1</sup>: Black dots, heating curve; red dots, cooling curve. (<b>a</b>) Polyester <b>3a</b>; (<b>b</b>) Polyester <b>3b</b>; and (<b>c</b>) Polyester <b>3c</b>.</p>
Full article ">Figure 2
<p>Release profiles of ketoprofen from polyester in different solution (pH 1.2, 7.4 and 10.0) at 37 °C. (<b>a</b>) Polyester <b>3a</b>; (<b>b</b>) Polyester <b>3b</b>; and (<b>c</b>) Polyester <b>3c</b>.</p>
Full article ">Figure 3
<p>Enzymatic copolymerization of polyethylene glycol (PEG) and 2-ketoprofen malic acid (thiomalic acid or aspartic acid) dimethyl ester.</p>
Full article ">
4177 KiB  
Review
Nanomembranes and Nanofibers from Biodegradable Conducting Polymers
by Elena Llorens, Elaine Armelin, María Del Mar Pérez-Madrigal, Luís Javier Del Valle, Carlos Alemán and Jordi Puiggalí
Polymers 2013, 5(3), 1115-1157; https://doi.org/10.3390/polym5031115 - 17 Sep 2013
Cited by 86 | Viewed by 23288
Abstract
This review provides a current status report of the field concerning preparation of fibrous mats based on biodegradable (e.g., aliphatic polyesters such as polylactide or polycaprolactone) and conducting polymers (e.g., polyaniline, polypirrole or polythiophenes). These materials have potential biomedical applications (e.g., tissue engineering [...] Read more.
This review provides a current status report of the field concerning preparation of fibrous mats based on biodegradable (e.g., aliphatic polyesters such as polylactide or polycaprolactone) and conducting polymers (e.g., polyaniline, polypirrole or polythiophenes). These materials have potential biomedical applications (e.g., tissue engineering or drug delivery systems) and can be combined to get free-standing nanomembranes and nanofibers that retain the better properties of their corresponding individual components. Systems based on biodegradable and conducting polymers constitute nowadays one of the most promising solutions to develop advanced materials enable to cover aspects like local stimulation of desired tissue, time controlled drug release and stimulation of either the proliferation or differentiation of various cell types. The first sections of the review are focused on a general overview of conducting and biodegradable polymers most usually employed and the explanation of the most suitable techniques for preparing nanofibers and nanomembranes (i.e., electrospinning and spin coating). Following sections are organized according to the base conducting polymer (e.g., Sections 4–6 describe hybrid systems having aniline, pyrrole and thiophene units, respectively). Each one of these sections includes specific subsections dealing with applications in a nanofiber or nanomembrane form. Finally, miscellaneous systems and concluding remarks are given in the two last sections. Full article
(This article belongs to the Special Issue Semiconducting Polymers for Organic Electronic Devices)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme showing the chemical structure of main biodegradable polymer used in biomedicine.</p>
Full article ">Figure 2
<p>Scheme showing the chemical structure of main conducting polymers.</p>
Full article ">Figure 3
<p>Schematic diagram showing the electrospinning process.</p>
Full article ">Figure 4
<p>Spin coating process involves the following steps: (<b>a</b>) Solution deposition; (<b>b</b>) Substrate acceleration; (<b>c</b>) Constant spinning rate; and (<b>d</b>) Drying and separation (not shown).</p>
Full article ">Figure 5
<p>Chemical structure of polyaniline (emeraldine base) and transformation to a conductive salt by prototonation in an acid medium.</p>
Full article ">Figure 6
<p>Scheme showing the preparation of self-assembled monolayers of ATQD-RGD, inset reproduced with permission from [<a href="#B41-polymers-05-01115" class="html-bibr">41</a>]. Copyright 2007 American Chemical Society.</p>
Full article ">Figure 7
<p>Synthesis scheme for the preparation of PGAT diblock copolymers.</p>
Full article ">Figure 8
<p>Atomic force microscopy (AFM) images of PLLA/PGAT samples containing (<b>a</b>) 33 wt % and (<b>b</b>) 10 wt % of PGAT. Reprinted with permission from [<a href="#B42-polymers-05-01115" class="html-bibr">42</a>]. Copyright 2011 WILEY-VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 9
<p>(<b>a</b>) Scheme based on [<a href="#B43-polymers-05-01115" class="html-bibr">43</a>] showing the synthesis of the multialdehide sodium alginate (MASA) and the tetraaniline-<span class="html-italic">graft</span>-multialdehide sodium alginate (MASA-AT); (<b>b</b>) Self-assembling and crosslinking capabilities of MASA-AT molecules; and (<b>c</b>) Conversions of MASA-AT between different oxidation states. Copyright 2011 WILEY-VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 10
<p>Two-step approach to prepare degradable and conductive block copolymers having aniline tetramer end groups.</p>
Full article ">Figure 11
<p>Structure and synthesis scheme for poly[(glycine ethyl ester)<span class="html-italic"><sub>x</sub></span>(aniline pentamer)<span class="html-italic"><sub>y</sub></span> phosphacene] (PGAP).</p>
Full article ">Figure 12
<p>Schematic synthesis route and structure of PLA-<span class="html-italic">co</span>-AP copolymer.</p>
Full article ">Figure 13
<p>Schematic representation of the self-assembling of PLA-<span class="html-italic">b</span>-AP-<span class="html-italic">b</span>-PLA triblock copolymers.</p>
Full article ">Figure 14
<p>Chemical structure of aniline pentamer cross-linking chitosan (AP-<span class="html-italic">cs</span>-CS) (<b>left</b>) and scanning micrographs (<b>right</b>) showing nanomicelles of AP-<span class="html-italic">cs</span>-CS and the typical morphology of CS. Reprinted with permission from [<a href="#B49-polymers-05-01115" class="html-bibr">49</a>]. Copyright 2008 American Chemical Society.</p>
Full article ">Figure 15
<p>Model explaining the higher conductivity of hyperbranched copolymers than the linear ones with the same content of conductive units.</p>
Full article ">Figure 16
<p>Chemical structure of copolymers constituted by anyline and aminobenzoic units.</p>
Full article ">Figure 17
<p>Scheme illustrating the key components of a novel biodegradable conducting polymer: conducting pyrrole-thiophene-pyrrole oligomer, degradable ester linkages and aliphatic linker.</p>
Full article ">Figure 18
<p>Chemical structure and synthesis of polycaprolactone fumarate (PCLF).</p>
Full article ">Figure 19
<p>AFM micrographs of PCLF-PPy (PCLT molecular weight of 18,000 g∙mol<sup>−1</sup>) surface microstructure (<b>left</b>) with root mean squared (RMS) roughness of 1195 nm and nanostructure with an RMS roughness of 8 nm (<b>right</b>). Reprinted with permission from [<a href="#B64-polymers-05-01115" class="html-bibr">64</a>]. Copyright 2010 Elsevier Ltd.</p>
Full article ">Figure 20
<p>SEM micrographs of (<b>a</b>) mesh-PPy scaffold; (<b>b</b>) Novel fluffy-PPy scaffold; (<b>c</b>) PPy coated PLLA fibers; and (<b>d</b>) TEM image of a PPy hollow fiber. Reprinted with permission from [<a href="#B87-polymers-05-01115" class="html-bibr">87</a>]. Copyright 2012 Royal Society of Chemistry.</p>
Full article ">Figure 21
<p>TEM images of (<b>a</b>,<b>b</b>) burr-shaped PPy hollow fibers polymerized under mechanical stirring and (<b>c</b>,<b>d</b>) PPy hollow fibers polymerized under ultrasonication. Reprinted with permission from [<a href="#B76-polymers-05-01115" class="html-bibr">76</a>]. Copyright 2012 Royal Society of Chemistry.</p>
Full article ">Figure 22
<p>Implant constituted by electrically actuable nanoporous membranes (based on [<a href="#B92-polymers-05-01115" class="html-bibr">92</a>]): nanoporous are closed at the reduction state due to PPy expansion whereas are open at the oxidation state making feasible the drug release.</p>
Full article ">Figure 23
<p>Formation of linear cell-seeded biosynthetic fiber constructs. (<b>a</b>) The two-step fabrication of the hybrid platform involved: (<b>1</b>) wet spinning of PLA:PLGA fibers onto a substrate to create an aligned microfiber array pattern; and (<b>2</b>) electrochemical deposition on the substrate of the conducting PPy; (<b>b</b>) The compatibility of the hybrid platform toward skeletal muscle was assessed by: (<b>1</b>) Proliferation and adhesion of cells and (<b>2</b>) Cell differentiation; and (<b>c</b>) The hybrid scaffold can be removed from the substrate layer and manipulated into 3D structures for <span class="html-italic">in vivo</span> implantation as required. Reprinted with permission from [<a href="#B100-polymers-05-01115" class="html-bibr">100</a>] and [<a href="#B102-polymers-05-01115" class="html-bibr">102</a>]. Copyright 2009 WILEY-VCH Verlag GmbH &amp; Co. and 2011 American Chemical Society.</p>
Full article ">Figure 24
<p>Synthesis of poly(5,5′′′-bis(hydroxymethyl)-3,3′′′-dimethyl-2,2′,5′,2′′,5′′,2′′′-quaterthiophene-<span class="html-italic">co</span>-Adipic Acid) (QAPE).</p>
Full article ">Figure 25
<p>Scanning electron micrographs of (<b>a</b>) PLGA nanosfibers and (<b>b</b>) poly(3,4-ethylenedioxythiophene) (PEDOT) nanotubes after removing the PLGA core fibers; Schematic illustrations of (<b>c</b>) the controlled release of dexamethasone and (<b>d</b>) the control of the release by applying an external electrical stimulation and positive charges in the polymer chains are compensated. To maintain overall charge neutrality, counterions are expelled towards the solution, nanotubes contract and drugs come out of the ends of tubes. Reprinted with permission from [<a href="#B116-polymers-05-01115" class="html-bibr">116</a>]. Copyright 2006 WILEY-VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 26
<p>Schematic illustration of conducting polymer (PEDOT) nanotube fabrication on neural microelectrodes: (<b>a</b>) Electrospinning of biodegradable PLGA template fibers with well-defined surface texture (<b>1</b>); (<b>b</b>) Electrochemical deposition of conducting polymer (PEDOT) around the electrospun fibers (<b>2</b>); (<b>c</b>) Dissolving the electrospun core fibers to create conducting polymer nanotubes (<b>3</b>); Optical microscopy images of (<b>d</b>) the entire microelectrode site before surface modification; The electrode site (<b>e</b>) after electrospinning of PLGA nanofibers; (<b>f</b>) After electrochemical deposition of PEDOT; and (<b>g</b>) After removing the PLGA core fibers. Reprinted with permission from [<a href="#B116-polymers-05-01115" class="html-bibr">116</a>]. Copyright 2006 WILEY-VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 27
<p>Schematic illustration showing the release of the anti-inflammatory drug from electrospun nanofibers (<b>a</b>) without and (<b>b</b>) with an alginate hydrogel coating; (<b>c</b>) Comparison of cumulative mass release profiles of DEX-loaded PLGA scaffolds with and without the alginate coating. Reprinted with permission from [<a href="#B125-polymers-05-01115" class="html-bibr">125</a>]. Copyright 2009 WILEY-VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 28
<p>Schemes showing (<b>left</b>) the coaxial electrospinning setup for P3HT electrospinning and (<b>right</b>) the elongation of P3HT domains and formation of continuous P3HT fibrils from the elongation of P3HT domains in highly concentrated PCL solutions.</p>
Full article ">Figure 29
<p>Schemes showing (<b>left</b>) the coaxial electrospinning setup for P3HT electrospinning and (<b>right</b>) the elongation of P3HT domains and formation of continuous P3HT fibrils from the elongation of P3HT domains in highly concentrated PCL solutions. Reprinted with permission from [<a href="#B128-polymers-05-01115" class="html-bibr">128</a>]. Copyright 2009 Royal Society of Chemistry.</p>
Full article ">Figure 30
<p>Chemical structures of poly(3-thiophene methyl acetate) (P3TMA) (<b>left</b>), and poly(tetramethylene succinate) (PE44) (<b>middle</b>) and digital camera image of a P3TMA:PE44 (<b>right</b>) free-standing nanomembrane dispersed in ethanol.</p>
Full article ">Figure 31
<p>(<b>a</b>) Enzymatic activity of lysozyme evaluated through degradation of the peptidoglycan of <span class="html-italic">Micrococcus luteus</span> (<span class="html-italic">i.e.</span>, decay on the 450 nm absorbance) of: PEDOT (control, curve 1); PEDOT coated with lysozyme (curve 2); PEDOT/lysozyme composite (curve 3) and free lysozyme (curve 4); (<b>b</b>) Control voltammograms of PEDOT (curve 1); PEDOT coate dwith lysozime (curve 2) and PEDOT/lysozime composite (curve 3).</p>
Full article ">Figure 32
<p>Chemical structure of the alginic acid-benzimidazole complex.</p>
Full article ">
1018 KiB  
Article
Hybrid, Nanoscale Phospholipid/Block Copolymer Vesicles
by Seng Koon Lim, Hans-Peter De Hoog, Atul N. Parikh, Madhavan Nallani and Bo Liedberg
Polymers 2013, 5(3), 1102-1114; https://doi.org/10.3390/polym5031102 - 6 Sep 2013
Cited by 62 | Viewed by 12842
Abstract
Hybrid phospholipid/block copolymer vesicles, in which the polymeric membrane is blended with phospholipids, display interesting self-assembly behavior, incorporating the robustness and chemical versatility of polymersomes with the softness and biocompatibility of liposomes. Such structures can be conveniently characterized by preparing giant unilamellar vesicles [...] Read more.
Hybrid phospholipid/block copolymer vesicles, in which the polymeric membrane is blended with phospholipids, display interesting self-assembly behavior, incorporating the robustness and chemical versatility of polymersomes with the softness and biocompatibility of liposomes. Such structures can be conveniently characterized by preparing giant unilamellar vesicles (GUVs) via electroformation. Here, we are interested in exploring the self-assembly and properties of the analogous nanoscale hybrid vesicles (ca. 100 nm in diameter) of the same composition prepared by film-hydration and extrusion. We show that the self-assembly and content-release behavior of nanoscale polybutadiene-b-poly(ethylene oxide) (PB-PEO)/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) hybrid phospholipid/block copolymer vesicles can be tuned by the mixing ratio of the amphiphiles. In brief, these hybrids may provide alternative tools for drug delivery purposes and molecular imaging/sensing applications and clearly open up new avenues for further investigation. Full article
(This article belongs to the Special Issue Supramolecular Chemistry and Self-Assembly 2013)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of nanoscale phospholipid/block copolymer hybrid vesicles.</p>
Full article ">Figure 2
<p>Flow-cytometric analysis of vesicle samples of (<b>a</b>) PB-PEO/POPC (75:25); (<b>b</b>) PB-PEO/POPC (50:50); (<b>c</b>) PB-PEO/POPC (25:75). The samples have been labeled with TMRho-PB-PEO and pyrene-PE to confirm the presence of hybrid vesicles in the population; (<b>d</b>) Control PB-PEO sample (unlabeled); (<b>e</b>) Control PB-PEO sample (pyrene-labeled); and (<b>f</b>) Control PB-PEO sample (TMRho-labeled). The <span class="html-italic">x</span> and <span class="html-italic">y</span> axes of each dot plot represent the fluorescence intensity of pyrene and TMRho, respectively.</p>
Full article ">Figure 3
<p>Hybrid giant unilamellar vesicles (GUVs) of PB-PEO/POPC (50:50) prepared by electroformation observed using fluorescence microscopy. Hybrid GUVs only form part of the GUV population. The green color indicates GUVs formed from PB-PEO, while red indicates POPC. Yellow indicates the presence of both, <span class="html-italic">i.e.</span>, hybrid vesicles.</p>
Full article ">Figure 4
<p>Intensity-weighted hydrodynamic diameter of extruded vesicles prepared from POPC, PB-PEO/POPC (75:25), PB-PEO/POPC (50:50), PB-PEO/POPC (25:75), and PB-PEO, as measured byDLS. All vesicles show a single unimodal population.</p>
Full article ">Figure 5
<p>The effect of PB-PEO molar ratio on the encapsulation efficiency of the vesicles. The hydrophilic dye carboxyfluorescein (CF) was used as model compound to calculate the encapsulation efficiency (% encapsulation) of vesicles. The encapsulation efficiency was calculated via 100 × <span class="html-italic">F</span><sub>t</sub>/<span class="html-italic">F</span><sub>0</sub>, where <span class="html-italic">F</span><sub>t</sub> is the encapsulated CF concentration (as calculated by CF fluorescence after 0.5% Triton X-100 addition) and <span class="html-italic">F</span><sub>0</sub> is the initial CF concentration of the stock solution.</p>
Full article ">Figure 6
<p>CF release from vesicles over 120 h at room temperature. 50 mM CF was encapsulated in the vesicles (<span class="html-italic">i.e.</span>, at self-quenching concentration). As a result, fluorescence intensity will increase when CF is released from the vesicles in the surrounding aqueous buffer. CF was excited at 480 nm and fluorescence emission at 520 nm was measured. The percentage of CF release over time is presented as 100 × (F − F<sub>0</sub>)/(F<sub>T</sub>− F<sub>0</sub>), where F<sub>0</sub> is the initial fluorescence of CF, F is the fluorescence of CF at time interval <span class="html-italic">t</span> and F<sub>T</sub> is the fluorescence intensity after complete release of CF after adding 0.5% Triton X-100.</p>
Full article ">
2122 KiB  
Article
Synthesis and Solution Properties of Double Hydrophilic Poly(ethylene oxide)-block-poly(2-ethyl-2-oxazoline) (PEO-b-PEtOx) Star Block Copolymers
by Tobias Rudolph, Sarah Crotty, Moritz Von der Lühe, David Pretzel, Ulrich S. Schubert and Felix H. Schacher
Polymers 2013, 5(3), 1081-1101; https://doi.org/10.3390/polym5031081 - 2 Sep 2013
Cited by 24 | Viewed by 9408
Abstract
We demonstrate the synthesis of star-shaped poly(ethylene oxide)-block-poly(2-ethyl-2-oxazoline) [PEOm-b-PEtOxn]x block copolymers with eight arms using two different approaches, either the “arm-first” or the “core-first” strategy. Different lengths of the outer PEtOx blocks ranging from [...] Read more.
We demonstrate the synthesis of star-shaped poly(ethylene oxide)-block-poly(2-ethyl-2-oxazoline) [PEOm-b-PEtOxn]x block copolymers with eight arms using two different approaches, either the “arm-first” or the “core-first” strategy. Different lengths of the outer PEtOx blocks ranging from 16 to 75 repeating units were used, and the obtained materials [PEO28-b-PEtOxx]8 were characterized via size exclusion chromatography (SEC), nuclear magnetic resonance spectroscopy (NMR), and Fourier-transform infrared spectroscopy (FT-IR) measurements. First investigations regarding the solution behavior in water as a non-selective solvent revealed significant differences. Whereas materials synthesized via the “core-first” method seemed to be well soluble (unimers), aggregation occurred in the case of materials synthesized by the “arm-first” method using copper-catalyzed azide-alkyne click chemistry. Full article
(This article belongs to the Special Issue Non-Equilibrium Blockcopolymer Self-Assembly)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>2D-LC results (acetonitrile (ACN)/H<sub>2</sub>O 55/45 v/v) for [PEO<sub>28</sub>-OH]<sub>8</sub> (<b>A</b>), [PEO<sub>28</sub>-Ts]<sub>8</sub> (<b>B</b>) and [PEO<sub>28</sub>-N<sub>3</sub>]<sub>8</sub> (<b>C</b>); in comparison with the SEC traces obtained for [PEO<sub>28</sub>-OH]<sub>8</sub> (solid black line), [PEO<sub>28</sub>-Ts]<sub>8</sub> (solid red line) and [PEO<sub>28</sub>-N<sub>3</sub>]<sub>8</sub> (solid blue line; THF was used as eluent).</p>
Full article ">Figure 2
<p>SEC traces using DMAC/LiCl as the eluent for TB-PEtOx<sub>x</sub> (dotted line), [PEO<sub>28</sub>-N<sub>3</sub>]<sub>8</sub> (scattered line) and the obtained purified star-shaped [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>x</sub>]<sub>8</sub> (solid line): (<b>A</b>) [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>18</sub>]<sub>8</sub>; (<b>B</b>) [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>55</sub>]<sub>8</sub>; (<b>C</b>) [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>75</sub>]<sub>8</sub>.</p>
Full article ">Figure 3
<p>First order time-conversion plot for the kinetic investigation of the microwave assisted polymerization of EtOx with [PEO<sub>28</sub>-Ts]<sub>8</sub> as the initiator at 140 °C (<b>A</b>); comparison of the time-dependent SEC traces (CHCl<sub>3</sub>) for the polymerization of EtOx (<b>B</b>).</p>
Full article ">Figure 4
<p>Time-dependent EtOx conversion (black squares) and the corresponding degrees of polymerization per arm (red squares) determined from the reaction mixture (filled squares) and after purification of the star-shaped block copolymers (empty squares) via NMR (<b>A</b>); SEC traces before (dashed line) and after purification via fractionated precipitation (<b>B</b>) (solid lines; CHCl<sub>3</sub> was used as the eluent).</p>
Full article ">Figure 5
<p>Comparison of the SEC traces obtained via the “core-first” (solid red lines) and the “arm-first” approach (solid black lines) in comparison to [PEO<sub>28</sub>-Ts]<sub>8</sub> (dashed line) for two compositions: (<b>A</b>) [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>16</sub>]<sub>8</sub> (core-first; red curve) and [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>18</sub>]<sub>8</sub> (arm-first; black curve); and (<b>B</b>) [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>50</sub>]<sub>8</sub> (core-first; red curve) and [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>55</sub>]<sub>8</sub> (arm-first; black curve).</p>
Full article ">Figure 6
<p>DLS CONTIN plot for “arm-first” approach stars in different solvents: in THF: [PEO<sub>28</sub>-OH]<sub>8</sub> (dashed black line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 1.5 nm), [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>18</sub>]<sub>8</sub> (red line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 2.5 nm), [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>55</sub>]<sub>8</sub> (green line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 4 nm) and [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>75</sub>]<sub>8</sub> (blue line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 5 nm) (2 g L<sup>-1</sup>) (<b>A</b>); in water [PEO<sub>28</sub>-OH]<sub>8</sub> (dotted black line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 3 nm), [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>18</sub>]<sub>8</sub> (red dashed, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 6 nm) and [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>75</sub>]<sub>8</sub> (blue line, &lt;R<sub>h</sub>&gt;<sub>n,app</sub> = 14 nm) (0.5 g L<sup>−1</sup>, filtered) (<b>B</b>).</p>
Full article ">Figure 7
<p>TEM micrographs of aggregates formed by [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>18</sub>]<sub>8</sub> (<b>top</b>) and [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>75</sub>]<sub>8</sub> (<b>bottom</b>) after heating for three days at 80 °C (unfiltered solutions, “arm-first” approach).</p>
Full article ">Figure 8
<p>Preparation of [PEO<sub>28</sub>-Ts]<sub>8</sub> and [PEO<sub>28</sub>-N<sub>3</sub>]<sub>8</sub>.</p>
Full article ">Figure 9
<p>Schematic representation of the synthesis of star-shaped [PEO<sub>28</sub>-<span class="html-italic">b</span>-PEtOx<sub>x</sub>]<sub>8</sub> block copolymers using copper-catalyzed azide-alkyne cycloaddition click chemistry (CuAAC).</p>
Full article ">
801 KiB  
Article
Electrochemical and Spectroelectrochemical Properties of a New Donor–Acceptor Polymer Containing 3,4-Dialkoxythiophene and 2,1,3-Benzothiadiazole Units
by Erika Herrera Calderon, Milind Dangate, Norberto Manfredi, Alessandro Abbotto, Matteo M. Salamone, Riccardo Ruffo and Claudio M. Mari
Polymers 2013, 5(3), 1068-1080; https://doi.org/10.3390/polym5031068 - 12 Aug 2013
Cited by 7 | Viewed by 8019
Abstract
A new heteroarylene-vinylene donor-acceptor low bandgap polymer, the poly(DEHT-V-BTD), containing vinylene-spaced efficient donor (dialkoxythiophene) and acceptor (benzothiadiazole) moieties, is presented. Electropolymerization has been carried out by several electrochemical techniques and the results are compared. In particular, the pulsed potentiostatic method was able to [...] Read more.
A new heteroarylene-vinylene donor-acceptor low bandgap polymer, the poly(DEHT-V-BTD), containing vinylene-spaced efficient donor (dialkoxythiophene) and acceptor (benzothiadiazole) moieties, is presented. Electropolymerization has been carried out by several electrochemical techniques and the results are compared. In particular, the pulsed potentiostatic method was able to provide layers with sufficient amounts of material. Cyclic voltammetries showed reversible behavior towards both p- and n-doping. The HOMO, LUMO, and bandgap energies were estimated to be −5.3, −3.6 and 1.8 eV, respectively. In situ UV-Vis measurements have established that the presence of the vinylene group stabilizes the formation of polaronic charge carriers even at high doping levels. Full article
(This article belongs to the Special Issue Semiconducting Polymers for Organic Electronic Devices)
Show Figures

Figure 1

Figure 1
<p>The monomer (DEHT-V-BTD).</p>
Full article ">Figure 2
<p>Synthesis of monomer DEHT-V-BTD.</p>
Full article ">Figure 3
<p>Absorption (continuous line) and emission spectra (dashed line) of monomer DEHT-V-BTD in CHCl<sub>3</sub>.</p>
Full article ">Figure 4
<p>Polymerization of DEHT-V-BTD by different electrochemical techniques: (<b>a</b>) three cyclic voltammetry (CV) cycles; (<b>b</b>) potentiostatic (PS) at 0.39 V for 100 s: current/time (black) and charge/time (red) profiles; (<b>c</b>) first six steps of the pulsed-potentiostatic (PPS): current/time (black) and potential/time (red) profiles; and (<b>d</b>) PPS: current/time (black) and charge/time (red) profiles.</p>
Full article ">Figure 5
<p>CVs (20 mV/s) of different poly(DEHT-V-BTD) in monomer free solution: sample obtained by: (<b>a</b>) CV; (<b>b</b>) PS; (<b>c</b>) and (<b>d</b>) PPS.</p>
Full article ">Figure 6
<p>Spectroelectrochemistry in different potential regions of a poly(DEHT-V-BTD) film prepared by PPS: (<b>a</b>) oxidation of the neutral polymer; (<b>b</b>) reduction of the oxidized polymer; (<b>c</b>) reduction of the neutral polymer; and (<b>d</b>) oxidation of the reduced polymer. OCP refers to the open circuit potential spectrum.</p>
Full article ">
229 KiB  
Article
Preparation and Characterization of Poly(ethyl hydrazide) Grafted Oil Palm Empty Fruit Bunch for Removal of Ni(II) Ion in Aqueous Environment
by Ili Syazana bt Johari, Nor Azah Yusof, Md Jelas Haron and Siti Mariam Mohd Nor
Polymers 2013, 5(3), 1056-1067; https://doi.org/10.3390/polym5031056 - 25 Jul 2013
Cited by 15 | Viewed by 6455
Abstract
Poly(ethyl hydrazide) grafted oil palm empty fruit bunch (peh-g-opefb) fiber has been successfully prepared by heating poly(methylacrylate)-g-opefb at 60 °C for 4 h in a solution of hydrazine hydrate in ethanol. The chelating ability of peh-g-opefb was evaluated based on removal of Ni(II) [...] Read more.
Poly(ethyl hydrazide) grafted oil palm empty fruit bunch (peh-g-opefb) fiber has been successfully prepared by heating poly(methylacrylate)-g-opefb at 60 °C for 4 h in a solution of hydrazine hydrate in ethanol. The chelating ability of peh-g-opefb was evaluated based on removal of Ni(II) ions in aqueous solution. Adsorption of Ni(II) by peh-g-opefb was characterized based on effect of pH, isotherm, kinetic and thermodynamic study. This cheap sorbent based on oil palm empty fruit bunch fiber has a great future potential in water treatment industries based on high adsorption capacity, biodegradability and renewability. Full article
Show Figures

Figure 1

Figure 1
<p>A complex of poly(acryloyl benzoic hydrazide) with nickel ion [<a href="#B13-polymers-05-01056" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>FTIR spectra of opefb, pma-g-opefb and peh-g-opefb.</p>
Full article ">Figure 3
<p>Adsorption capacity of peh-g-opefb towards nickel ion at different pH.</p>
Full article ">Figure 4
<p>Effect of initial concentration of nickel ions on adsorption capacities by peh-g-opefb at different temperature.</p>
Full article ">Figure 5
<p>Adsorption capacity towards Ni(II) ion at different times.</p>
Full article ">
755 KiB  
Article
DNA-Promoted Auto-Assembly of Gold Nanoparticles: Effect of the DNA Sequence on the Stability of the Assemblies
by Matthieu Doyen, Kristin Bartik and Gilles Bruylants
Polymers 2013, 5(3), 1041-1055; https://doi.org/10.3390/polym5031041 - 22 Jul 2013
Cited by 7 | Viewed by 7309
Abstract
The use of deoxyribonucleic acid (DNA) oligonucleotides has proven to be a powerful and versatile strategy to assemble nanomaterials into two (2D) and three-dimensional (3D) superlattices. With the aim of contributing to the elucidation of the factors that affect the stability of this [...] Read more.
The use of deoxyribonucleic acid (DNA) oligonucleotides has proven to be a powerful and versatile strategy to assemble nanomaterials into two (2D) and three-dimensional (3D) superlattices. With the aim of contributing to the elucidation of the factors that affect the stability of this type of superlattices, the assembly of gold nanoparticles grafted with different DNA oligonucleotides was characterized by UV-Vis absorption spectroscopy as a function of temperature. After establishing an appropriate methodology the effect of (i) the length of the grafted oligonucleotides; (ii) the length of their complementary parts and also of (iii) the simultaneous grafting of different oligonucleotides was investigated. Our results indicate that the electrostatic repulsion between the particles and the cooperativity of the assembly process play crucial roles in the stability of the assemblies while the grafting density of the oligonucleotide strands seems to have little influence. Full article
(This article belongs to the Special Issue Supramolecular Chemistry and Self-Assembly 2013)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the assembly of two sets of nanoparticles functionalized with oligonucleotides with complementary segments.</p>
Full article ">Figure 2
<p>Absorbance at 260 nm as a function of temperature of a solution of AuNP-DNAA<sub>11</sub> mixed with AuNP-DNAT<sub>11</sub>. Three assembly/disassembly cycles are displayed.</p>
Full article ">Figure 3
<p>Absorbance at 260 nm as a function of temperature during assembly and disassembly cycle of complementary AuNP-DNA (AuNP-T10-DNAA<sub>11</sub> mixed with AuNP-T10-DNAT<sub>11</sub>) recorded using three different heating/cooling rates: 1 °C/min (dashed line); 0.5 °C/min (solid line); 0.1 °C/min (bold line).</p>
Full article ">Figure 4
<p>Normalized absorbance at 260 nm during assembly of AuNP functionalized with T10-DNAT<sub>11</sub> and mixed AuNP functionalized with (i) T10-DNAA<sub>11</sub> (plain); (ii) T10-DNAA<sub>11</sub>/T10-DNAA<sub>5</sub> (25/75) (dotted); (iii) T10-DNAA<sub>11</sub>/T10-DNAA<sub>5</sub> (50/50) (dotted-point); (iv) T10-DNAA<sub>11</sub>/T10-DNAA<sub>5</sub> (75/25) (dotted-double point) and (v) T10-DNAA<sub>5</sub> (bold) in function of temperature.</p>
Full article ">Figure 5
<p>Absorption spectra during the cooling ramp of the assembly AuNPs-DNAADNAT<sub>11</sub> mixed with AuNPs-DNAT<sub>11</sub>.</p>
Full article ">Figure 6
<p>Normalized absorbance at 260 nm during assembly of DNA-AuNPs functionalized with the complementary oligonucleotides DNAA<sub>11</sub> &amp; DNAT<sub>11</sub> (plain), DNAA<sub>11</sub> &amp; T10-DNAT<sub>11</sub> (dotted) and T10- DNAA<sub>11</sub> &amp; T10- DNAT<sub>11</sub> (bold) as a function of temperature.</p>
Full article ">Figure 7
<p>Thermal denaturation curve of the DNAA11-DNAT11 duplex (5 µM) followed by UV-Vis absorption spectroscopy at 260 nm in a 10 mM phosphate buffer containing 500 mM NaCl at pH = 7.4.</p>
Full article ">Figure 8
<p>Mean value of melting temperature and a confidence interval of 95% in a 100 mM NaCl(diamond), melting temperature in a 500 mM NaCl (square), predicted melting temperature in a 100 mM NaCl (plain) and predicted melting temperature in a 500 mM (bold) of the pairing between strands A<sub>11</sub> &amp; T<sub>11</sub> in function of duplex concentration. Prediction were made with <span class="html-italic">OligoCalc</span> [<a href="#B46-polymers-05-01041" class="html-bibr">46</a>].</p>
Full article ">Figure 9
<p>Normalized absorbance at 260 nm during the assembly of DNA-AuNPs functionalized with the complementary strands T10-DNAA<sub>11</sub> &amp; T10-DNAT<sub>11</sub> (plain), T10 DNAA<sub>11</sub> &amp; T10-DNAT<sub>5</sub> (dotted), and T10 DNAA<sub>5</sub> &amp; T10-DNAT<sub>11</sub> (bold) as a function of temperature.</p>
Full article ">
751 KiB  
Review
Precise Synthesis of Block Polymers Composed of Three or More Blocks by Specially Designed Linking Methodologies in Conjunction with Living Anionic Polymerization System
by Yuri Matsuo, Ryuji Konno, Takashi Ishizone, Raita Goseki and Akira Hirao
Polymers 2013, 5(3), 1012-1040; https://doi.org/10.3390/polym5031012 - 17 Jul 2013
Cited by 50 | Viewed by 14872
Abstract
This article reviews the successful development of two specially designed linking methodologies in conjunction with a living anionic polymerization system for the synthesis of novel multiblock polymers, composed of three or more blocks, difficult to be synthesized by sequential polymerization. The first methodology [...] Read more.
This article reviews the successful development of two specially designed linking methodologies in conjunction with a living anionic polymerization system for the synthesis of novel multiblock polymers, composed of three or more blocks, difficult to be synthesized by sequential polymerization. The first methodology with the use of a new heterofunctional linking agent, 2-(4-chloromethylphenyl)ethyldimethylchlorosilane (1), was developed for the synthesis of multiblock polymers containing poly(dimethylsiloxane) (PDMS) blocks. This methodology is based on the selective reaction of the chain-end silanolate anion of living PDMS, with the silyl chloride function of 1, and subsequent linking reaction of the resulting ω-chain-end-benzyl chloride-functionalized polymer with either a living anionic polymer or living anionic block copolymer. With this methodology, various multiblock polymers containing PDMS blocks, up to the pentablock quintopolymer, were successfully synthesized. The second methodology using an α-phenylacrylate (PA) reaction site was developed for the synthesis of multiblock polymers composed of all-vinyl polymer blocks. In this methodology, an α-chain-end-PA-functionalized polymer or block copolymer, via the living anionic polymerization, was first prepared and, then, reacted with appropriate living anionic polymer or block copolymer to link the two polymer chains. As a result, ACB (BCA), BAC (CAB), (AB)n, (AC)n, ABA, ACA, BCB, and ABCA multiblock polymers, where A, B, and C were polystyrene, poly(2-vinylpyridine), and poly(methyl methacrylate) segments, could be successfully synthesized. The synthesis of triblock copolymers, BAB, CAC, and CBC, having molecular asymmetry in both side blocks, was also achieved. Furthermore, the use of living anionic polymers, derived from many other monomers, categorized as either of styrene, 2-vinylpyridine, or methyl methacrylate in monomer reactivity, in the linking methodology enabled the number of synthetically possible block polymers to be greatly increased. Once again, all of the block polymers synthesized by these methodologies are new and cannot be synthesized at all by sequential polymerization. They were well-defined in block architecture and precisely controlled in block segment. Full article
(This article belongs to the Special Issue Non-Equilibrium Blockcopolymer Self-Assembly)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEC profiles for a series of multiblock copolymers: (<b>a</b>) AB diblock; (<b>b</b>) (AB)<sub>2</sub> tetrablock; (<b>c</b>) (AB)<sub>3</sub> hexablock; (<b>d</b>) (AB)<sub>4</sub> octablock; and (<b>e</b>) (AB)<sub>5</sub> decablock copolymers.</p>
Full article ">Figure 2
<p>Synthesis of PS-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-P2VP-<span class="html-italic">b</span>-PEO tetrablock quaterpolymer and PS-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-P2VP-<span class="html-italic">b</span>-P<sup>t</sup>BMA-<span class="html-italic">b</span>-PEO pentablock quintopolymer.</p>
Full article ">Figure 3
<p>Synthesis of PHIC-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-PS-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-PHIC pentablock terpolymer.</p>
Full article ">Figure 4
<p>Synthesis of P<sup>t</sup>BMA-<span class="html-italic">b</span>-PDMS and P2VP-<span class="html-italic">b</span>-PDMS diblock copolymers.</p>
Full article ">Figure 5
<p>Synthesis of PS-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-PDMS-<span class="html-italic">b</span>-P2VP tetrablock quaterpolymer.</p>
Full article ">Figure 6
<p>Synthesis of PS-<span class="html-italic">b</span>-PI-<span class="html-italic">b</span>-PDMS-<span class="html-italic">b</span>-P<sup>t</sup>BMA-<span class="html-italic">b</span>-P2VP pentablock quintopolymer.</p>
Full article ">Figure 7
<p>Synthesis of (<b>a</b>) PS-<span class="html-italic">b</span>-PMMA-<span class="html-italic">b</span>-PS and (<b>b</b>) PMMA-<span class="html-italic">b</span>-PS-<span class="html-italic">b</span>-PMMA triblock copolymers.</p>
Full article ">Figure 8
<p>Synthesis of (PS-<span class="html-italic">b</span>-PMMA)<span class="html-italic"><sub>n</sub></span> alternate multiblock copolymers.</p>
Full article ">Figure 9
<p>Synthesis of (<b>a</b>) PS-<span class="html-italic">b</span>-PMMA-<span class="html-italic">b</span>-P2VP and (<b>b</b>) P2VP-<span class="html-italic">b</span>-PS-<span class="html-italic">b</span>-PMMA triblock terpolymers.</p>
Full article ">Figure 10
<p>Deprotection of (<b>a</b>) P(Si-HEMA); (<b>b</b>) P(acetal-DIMA); and (<b>c</b>) P<sup>t</sup>BMA blocks.</p>
Full article ">Figure 11
<p>Synthesis of PS-<span class="html-italic">b</span>-P2VP-<span class="html-italic">b</span>-PMMA-<span class="html-italic">b</span>-PS tetrablock terpolymer.</p>
Full article ">Figure 12
<p>Synthesis of a rod-coil P3HP-<span class="html-italic">b</span>-PMMA diblock copolymer.</p>
Full article ">Figure 13
<p>Monomers, <b>a</b>, <b>b</b>, and <b>c</b> with different reactivities.</p>
Full article ">
1138 KiB  
Review
Design Strategies for Functionalized Poly(2-oxazoline)s and Derived Materials
by Elisabeth Rossegger, Verena Schenk and Frank Wiesbrock
Polymers 2013, 5(3), 956-1011; https://doi.org/10.3390/polym5030956 - 15 Jul 2013
Cited by 130 | Viewed by 24989
Abstract
The polymer class of poly(2-oxazoline)s currently is under intensive investigation due to the versatile properties that can be tailor-made by the variation and manipulation of the functional groups they bear. In particular their utilization in the biomedic(in)al field is the subject of numerous [...] Read more.
The polymer class of poly(2-oxazoline)s currently is under intensive investigation due to the versatile properties that can be tailor-made by the variation and manipulation of the functional groups they bear. In particular their utilization in the biomedic(in)al field is the subject of numerous studies. Given the mechanism of the cationic ring-opening polymerization, a plethora of synthetic strategies exists for the preparation of poly(2-oxazoline)s with dedicated functionality patterns, comprising among others the functionalization by telechelic end-groups, the incorporation of substituted monomers into (co)poly(2-oxazoline)s, and polymeranalogous reactions. This review summarizes the current state-of-the-art of poly(2-oxazoline) preparation and showcases prominent examples of poly(2-oxazoline)-based materials, which are retraced to the desktop-planned synthetic strategy and the variability of their properties for dedicated applications. Full article
(This article belongs to the Special Issue Ring-Opening Polymerization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images of a SiO<sub>2</sub>/SiOH-based tip before (left) and after (right) grafting of <b>pMeOx</b> on the surface (for details, see reference [<a href="#B100-polymers-05-00956" class="html-bibr">100</a>]).</p>
Full article ">Figure 2
<p>Reaction scheme for the methyl tosylate-initiated cationic ring-opening polymerization CROP of 2-oxazolines for the example of the block copolymerization of 2-methyl-2-oxazoline and 2-ethyl-2-oxazoline, yielding the diblock copolymer poly(2-methyl-2-oxazoline)-<span class="html-italic">block</span>-poly(2-ethyl-2-oxazoline) after termination with water (top). Poly(2-oxzoline)s can be subjected to acid-mediated partial hydrolysis [shown for the example of poly(2-ethyl-2-oxazoline)], yielding the random copolymer poly(2-ethyl-2-oxazoline)-<span class="html-italic">stat</span>-poly(ethylene imine) (bottom) and paving the way to further polymeranalogous reactions.</p>
Full article ">Figure 3
<p>Structural formula of telechelic <b>pMeOx</b> exhibiting antimicrobial activity (for details, see reference [<a href="#B44-polymers-05-00956" class="html-bibr">44</a>]).</p>
Full article ">Figure 4
<p>Structural formula of telechelic <b>(MeO-EO<sub>1</sub>)<span class="html-italic"><sup>n</sup></span>PrOx</b> mimicking a lipopolymer structure (for details, see reference [<a href="#B56-polymers-05-00956" class="html-bibr">56</a>]).</p>
Full article ">Figure 5
<p>Scheme of the reaction of a carboxylic acid-end-capped telechelic (bio)polyester with <b>OxOx</b>.</p>
Full article ">Figure 6
<p>Reaction scheme for the synthesis of <b>poly(</b><b><span class="html-small-caps">l</span>-lactic acid)-<span class="html-italic">block</span>-pEtOx-<span class="html-italic">block</span>-poly(</b><b><span class="html-small-caps">l</span>-lactic acid)</b> triblock copolymers from telechelic <b>pEtOx</b> macroinitiators, derived from the CROP of <b>EtOx</b> initiated with bisfunctional 1,4-dibromo-2-butene (for details, see reference [<a href="#B69-polymers-05-00956" class="html-bibr">69</a>]).</p>
Full article ">Figure 7
<p>Synthesis of the diblock copolymer <b>p<span class="html-italic"><sup>i</sup></span>PrOx-<span class="html-italic">block</span>-poly(</b><b>L</b><b>-glutamate)</b> involving amine-functionalized <b>p<span class="html-italic"><sup>i</sup></span>PrOx</b> macroinitiators and <span class="html-italic">N</span>-carboxyanhydride protected <span class="html-small-caps">l</span>-glutamic acid (for details, see reference [<a href="#B73-polymers-05-00956" class="html-bibr">73</a>]).</p>
Full article ">Figure 8
<p>Diblock copolymers formed by the supramolecular assembly of β-cyclodextrin-functionalized poly(<span class="html-italic">N</span>-isopropylacrylamide) and adamantine-end-capped <b>pMeOx</b> (for details, see reference [<a href="#B81-polymers-05-00956" class="html-bibr">81</a>]).</p>
Full article ">Figure 9
<p>Performance of the living anionic <span class="html-italic"><sup>n</sup></span>BuLi-initiated polymerization of <b><span class="html-italic"><sup>i</sup></span>Pr<sup>=</sup>Ox</b> and subsequent graft polymerization of 2-oxazolines from the pending oxazoline moieties (top; for details, see reference [<a href="#B92-polymers-05-00956" class="html-bibr">92</a>]). Grafting of poly(2-oxazoline)s from dedicatedly equipped silicon surfaces (for details, see reference [<a href="#B98-polymers-05-00956" class="html-bibr">98</a>]).</p>
Full article ">Figure 10
<p>Structural formulae of a hexatosylate initiator and the corresponding star-shaped <b>pEtOx</b> (for details, see reference [<a href="#B118-polymers-05-00956" class="html-bibr">118</a>]).</p>
Full article ">Figure 11
<p>Reaction scheme for the usage of 2nd generation propylene imine dendrimers as terminating agents during the CROP of <b>EtOx </b>(for details, see reference [<a href="#B125-polymers-05-00956" class="html-bibr">125</a>]).</p>
Full article ">Figure 12
<p>Three common strategies for the synthesis of 2-oxazoline monomers: from nitriles according to Witte and Seeliger (top), from carboxylic acids and esters according to the Henkel patent (middle), and the three-step synthesis from carboxylic acids involving activation by <span class="html-italic">N</span>-hydroxysuccinimide.</p>
Full article ">Figure 13
<p>Structural formulae of poly(2-oxazoline)s with chiral carbon atoms in the main chain: <b>p(<span class="html-italic">R</span>)-<span class="html-italic"><sup>n</sup></span>BuOx(Et)</b>, <b>p(<span class="html-italic">S</span>)-<span class="html-italic"><sup>n</sup></span>BuOx(Et)</b> and <b>p(<span class="html-italic">RS</span>)-<span class="html-italic"><sup>n</sup></span>BuOx(Et)</b> (top; for details, see reference [<a href="#B155-polymers-05-00956" class="html-bibr">155</a>]). Anionic polymerization of maleimides bearing optically active 2-oxazoline substituents (bottom; for details, see reference [<a href="#B151-polymers-05-00956" class="html-bibr">151</a>]).</p>
Full article ">Figure 14
<p>Structural formula of amphiphilic <b>pEtOx-<span class="html-italic">block</span>-pNonOx</b>.</p>
Full article ">Figure 15
<p>Structural formula of imidazole-functionalized poly(2-oxazoline)s as immobilized catalysts for Heck and Suzuki coupling reactions (top; for details, see reference [<a href="#B181-polymers-05-00956" class="html-bibr">181</a>]). Structural formula of bipyridine-functionalized poly(2-oxazoline)s that can be used as macroligands in atom-transfer radical polymerizations (bottom; for details, see reference [<a href="#B172-polymers-05-00956" class="html-bibr">172</a>]).</p>
Full article ">Figure 16
<p>Reaction scheme for the covalent attachment of 2,3,4,6-tetra-<span class="html-italic">O</span>-acetyl-1-thio-glucopyranose to <b>pDec<sup>=</sup>Ox</b> by thiol-ene reactions (top; for details, see reference [<a href="#B183-polymers-05-00956" class="html-bibr">183</a>]). Reaction scheme of the Huisgen cycloaddition of <b>pMeOx-<span class="html-italic">stat</span>-p<span class="html-italic"><sup>n</sup></span>Pe</b><b><sup>≡</sup></b><b>Ox</b> and small-molecule azides (bottom; for details, see reference [<a href="#B194-polymers-05-00956" class="html-bibr">194</a>]).</p>
Full article ">Figure 17
<p>Reaction scheme for the synthesis of multiple-generation linear dendrimers with a <b>pEI</b> core by sequential (repeated) reaction with methyl acrylate and ethylene diamine (for details, see reference [<a href="#B215-polymers-05-00956" class="html-bibr">215</a>]).</p>
Full article ">Figure 18
<p>Reaction scheme for the fabrication of functionalized surfaces: hydrolysis of <b>pEtOx</b>, reaction of the recovered <b>pEI</b> with 1-bromododecane and a brominated benzophenone linker, and subsequent C–H bond insertion (for details, see reference [<a href="#B217-polymers-05-00956" class="html-bibr">217</a>]).</p>
Full article ">
114 KiB  
Editorial
Open Access Makes an Impact
by Alexander Böker
Polymers 2013, 5(3), 954-955; https://doi.org/10.3390/polym5030954 - 5 Jul 2013
Viewed by 4941
Abstract
Polymers published its first issue in December 2009. At that time, the editorial board and publisher were determined to lead the journal to become another MDPI success story, proving that open access publishing and high quality publications, ensured by a rigorous peer-review procedure, [...] Read more.
Polymers published its first issue in December 2009. At that time, the editorial board and publisher were determined to lead the journal to become another MDPI success story, proving that open access publishing and high quality publications, ensured by a rigorous peer-review procedure, followed by fast publication of accepted manuscripts can be achieved. Three and a half years later, after more than 153,000 article downloads in 2012, the journal received its first impact factor (2012 JCR IF: 1.687). Today, Polymers is proud to be the number one open access journal in the category of “Polymer Science”. In order to achieve this, we relied on an editorial board with well-known members from the polymer community, a professional staff and a vision that there is room alongside the established “high impact” journals for publishing science. Recently, the editor in chief of Science, Bruce Alberts, wrote an editorial [1] in favor of the San Francisco Declaration on Research Assessment (DORA), which states that the impact factor must not be used as “a surrogate measure of the quality of individual research articles [2]”. Even though the impact factor certainly is so far the best available measure of the quality of a journal and its impact on the scientific community, when it comes to the single manuscripts published, “there is still no other way to evaluate the quality of scientific papers, but to read them [3]”. Therefore, Polymers celebrates its first impact factor with an appropriate critical distance towards bibliometric data and feels encouraged to continue on the chosen path. With this in mind, I encourage you, our readers, to continuously evaluate the scientific quality of the articles published in Polymers by reading, discussing and citing them. [...] Full article
366 KiB  
Article
Supramolecular Functionalities Influence the Thermal Properties, Interactions and Conductivity Behavior of Poly(ethylene glycol)/LiAsF6 Blends
by Jui-Hsu Wang, Chih-Chia Cheng, Oleksii Altukhov, Feng-Chih Chang and Shiao-Wei Kuo
Polymers 2013, 5(3), 937-953; https://doi.org/10.3390/polym5030937 - 4 Jul 2013
Cited by 7 | Viewed by 6980
Abstract
In this study, we tethered terminal uracil groups onto short-chain poly(ethylene glycol) (PEG) to form the polymers, uracil (U)-PEG and U-PEG-U. Through AC impedance measurements, we found that the conductivities of these polymers increased upon increasing the content of the lithium salt, LiAsF [...] Read more.
In this study, we tethered terminal uracil groups onto short-chain poly(ethylene glycol) (PEG) to form the polymers, uracil (U)-PEG and U-PEG-U. Through AC impedance measurements, we found that the conductivities of these polymers increased upon increasing the content of the lithium salt, LiAsF6, until the Li-to-PEG ratio reached 1:4, with the conductivities of the LiAsF6/U-PEG blends being greater than those of the LiAsF6/U-PEG-U blends. The ionic conductivity of the LiAsF6/U-PEG system reached as high as 7.81 × 10−4 S/cm at 30 °C. Differential scanning calorimetry, wide-angle X-ray scattering, 7Li nuclear magnetic resonance spectroscopy and Fourier transform infrared spectroscopy revealed that the presence of the uracil groups in the solid state electrolytes had a critical role in tuning the glass transition temperatures and facilitating the transfer of Li+ ions. Full article
(This article belongs to the Special Issue Supramolecular Chemistry and Self-Assembly 2013)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Differential scanning calorimetry (DSC) thermograms of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends.</p>
Full article ">Figure 2
<p>Values of <span class="html-italic">T<sub>g</sub></span> for U-PEG and U-PEG-U blends incorporating various lithium salt contents. Solid lines were calculated from the configuration entropy model.</p>
Full article ">Figure 3
<p>WAXD spectra of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends at room temperature.</p>
Full article ">Figure 4
<p>FTIR spectra (from 750 to 650 cm<sup>−1</sup>) of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends at room temperature.</p>
Full article ">Figure 5
<p>FTIR spectra (from 1200 to 1000 cm<sup>−1</sup>) of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends at room temperature.</p>
Full article ">Figure 6
<p>FTIR spectra (from 1800 to 1600 cm<sup>−1</sup>) of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends at room temperature.</p>
Full article ">Figure 7
<p>Solid state <sup>7</sup>Li proton-decoupled MAS NMR spectra of (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends at room temperature.</p>
Full article ">Figure 8
<p>Ionic conductivity plotted with respect to temperature for (<b>a</b>) LiAsF<sub>6</sub>/U-PEG and (<b>b</b>) LiAsF<sub>6</sub>/U-PEG-U blends.</p>
Full article ">Figure 9
<p>Chemical Structures of uracil (U)-poly(ethylene glycol) (PEG) and U-PEG-U.</p>
Full article ">Figure 10
<p>Peak assignments in the spectra of (<b>a</b>) U-PEG; (<b>b</b>) Li/U-PEG blends; (<b>c</b>) U-PEG-U; (<b>d</b>) Li/U-PEG-U blends.</p>
Full article ">
13847 KiB  
Article
Coarse-Grained Models for Protein-Cell Membrane Interactions
by Ryan Bradley and Ravi Radhakrishnan
Polymers 2013, 5(3), 890-936; https://doi.org/10.3390/polym5030890 - 2 Jul 2013
Cited by 46 | Viewed by 17020
Abstract
The physiological properties of biological soft matter are the product of collective interactions, which span many time and length scales. Recent computational modeling efforts have helped illuminate experiments that characterize the ways in which proteins modulate membrane physics. Linking these models across time [...] Read more.
The physiological properties of biological soft matter are the product of collective interactions, which span many time and length scales. Recent computational modeling efforts have helped illuminate experiments that characterize the ways in which proteins modulate membrane physics. Linking these models across time and length scales in a multiscale model explains how atomistic information propagates to larger scales. This paper reviews continuum modeling and coarse-grained molecular dynamics methods, which connect atomistic simulations and single-molecule experiments with the observed microscopic or mesoscale properties of soft-matter systems essential to our understanding of cells, particularly those involved in sculpting and remodeling cell membranes. Full article
(This article belongs to the Special Issue Multiscale Simulations in Soft Matter)
Show Figures

Figure 1

Figure 1
<p>Diagram of computational methods for studying biophysical systems across a range of time- and length-scales. Representative snapshots depict an all-atom lipid bilayer, peptides embedded in a coarse-grained bilayer and proteins remodeling a continuum mechanics membrane model. Bilayers were simulated with the CHARMM36 [<a href="#B15-polymers-05-00890" class="html-bibr">15</a>] and Martini [<a href="#B16-polymers-05-00890" class="html-bibr">16</a>] force fields and rendered with Visual Molecular Dynamics [<a href="#B17-polymers-05-00890" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>Representative snapshots of all-atom (upper right) and Martini coarse-grained (bottom) molecular dynamics simulations of a 4:1 dioleoylphosphatidylcholine with dioleoylphospatidylserine (DOPC/DOPS) bilayer. The upper left shows the coarse-grained mapping of a single DOPC lipid, with beads colored by bead type (gray for hydrocarbons-, pink for glycerol-, brown for phosphate- and blue for choline-type). The all-atom system contains 800 lipids, while the coarse-grained system contains 3,200 lipids (water molecules are not pictured here). Bilayers were simulated with the CHARMM36 [<a href="#B15-polymers-05-00890" class="html-bibr">15</a>] and Martini [<a href="#B16-polymers-05-00890" class="html-bibr">16</a>] force fields and rendered with Visual Molecular Dynamics [<a href="#B17-polymers-05-00890" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>Coarse-grained representation of the Martini model extension to amino acids [<a href="#B181-polymers-05-00890" class="html-bibr">181</a>], colored by bead type (where purple is apolar, blue and green are intermediate, gray and orange are polar and red represents charged particles).</p>
Full article ">Figure 4
<p>An example protein helix in all-atom (left) and Martini coarse-grained representations (center, backbone beads in gray and side-chain beads in yellow) with both images merged (right) to show how the fine-grained structure is mapped onto the coarse-grained beads. This image was rendered with Visual Molecular Dynamics [<a href="#B17-polymers-05-00890" class="html-bibr">17</a>].</p>
Full article ">Figure 5
<p>Simulations of six Bin/Amphiphysin/Rvs (BAR) domains remodeling a membrane. These simulations show that the proteins require a staggered arrangement (right) to bend the membrane, while the non-staggered arrangement (left) fails to generate curvature. Figure adapted from Arkhipov, <span class="html-italic">et al.</span> [<a href="#B105-polymers-05-00890" class="html-bibr">105</a>].</p>
Full article ">
3166 KiB  
Article
Extensional Flow Properties of Externally Plasticized Cellulose Acetate: Influence of Plasticizer Content
by Stefan Zepnik, Stephan Kabasci, Rodion Kopitzky, Hans-Joachim Radusch and Thomas Wodke
Polymers 2013, 5(3), 873-889; https://doi.org/10.3390/polym5030873 - 2 Jul 2013
Cited by 34 | Viewed by 11882
Abstract
Elongational flow properties of polymer melts are very important for numerous polymer processing technologies such as blown film extrusion or foam extrusion. Rheotens tests were conducted to investigate the influence of plasticizer content on elongational flow properties of cellulose acetate (CA). Triethyl citrate [...] Read more.
Elongational flow properties of polymer melts are very important for numerous polymer processing technologies such as blown film extrusion or foam extrusion. Rheotens tests were conducted to investigate the influence of plasticizer content on elongational flow properties of cellulose acetate (CA). Triethyl citrate (TEC) was used as plasticizer. Melt strength decreases whereas melt extensibility increases with increasing plasticizer content. Melt strength was further studied as a function of zero shear viscosity. The typical draw resonance of the Rheotens curve shifts to higher drawdown velocity and the amplitude of the draw resonance decreases with increasing TEC content. With respect to foam extrusion, not only are melt strength and melt extensibility important but the elongational behavior at low strain rates and the area under the Rheotens curve are also significant. Therefore, elongational viscosity as well as specific energy input were calculated and investigated with respect to plasticizer content. Preliminary foam extrusion tests of externally plasticized CA using chemical blowing agents confirm the results from rheological characterization. Full article
(This article belongs to the Collection Polysaccharides)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General structure of cellulose acetate (CA).</p>
Full article ">Figure 2
<p>General structure of triethyl citrate (TEC).</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic setup of the Rheotens test; (<b>b</b>) Typical Rheotens curve of externally plasticized CA: Drawdown force <span class="html-italic">F</span> as a function of drawdown velocity <span class="html-italic">v</span>.</p>
Full article ">Figure 4
<p>Torque and die pressure <span class="html-italic">p</span><sub>die</sub> as a function of extrusion time for CA/TEC (75/25) and 1% Unicell TS.</p>
Full article ">Figure 5
<p>(<b>a</b>) Die pressure <span class="html-italic">p</span><sub>die</sub> as a function of TEC content; (<b>b</b>) Specific mechanical energy input <span class="html-italic">SME</span> as a function of TEC content.</p>
Full article ">Figure 6
<p>(<b>a</b>) Shear viscosity <span class="html-italic">η</span> of externally plasticized CA as a function of shear rate <span class="html-fig-inline" id="polymers-05-00873-i004"> <img alt="Polymers 05 00873 i004" src="/polymers/polymers-05-00873/article_deploy/html/images/polymers-05-00873-i004.png"/></span> in dependence of TEC content; (<b>b</b>) Calculated master curve using a concentration-dependent shift factor <span class="html-italic">a</span><sub>c</sub> and predicted master curve using the Carreau-Yasuda model.</p>
Full article ">Figure 7
<p>(<b>a</b>) Drawdown force <span class="html-italic">F</span> as a function of draw ratio <span class="html-italic">V</span> in dependence of TEC content; (<b>b</b>) Mean drawdown force <span class="html-italic">F</span><sub>mid</sub> and maximum drawdown velocity <span class="html-italic">v</span><sub>max</sub> as function of TEC content.</p>
Full article ">Figure 8
<p>Apparent elongational viscosity of externally plasticized CA as a function of extensional rate calculated using the analytical model developed by Wagner <span class="html-italic">et al.</span> [<a href="#B31-polymers-05-00873" class="html-bibr">31</a>] in comparison to shear viscosity at low shear rate.</p>
Full article ">Figure 9
<p>(<b>a</b>) Influence of TEC content on the stress curves <span class="html-italic">σ</span> as a function of draw ratio <span class="html-italic">V</span>; (<b>b</b>) Initial slope of the stress curves <span class="html-italic">σ</span> via mean drawdown force <span class="html-italic">F</span><sub>mid</sub>.</p>
Full article ">Figure 10
<p>Maximum drawdown force <span class="html-italic">F</span><sub>max</sub> and mean drawdown force <span class="html-italic">F</span><sub>mid</sub> of externally plasticized CA via zero shear viscosity <span class="html-italic">η</span><sub>0</sub>.</p>
Full article ">Figure 11
<p>(<b>a</b>) Normalized drawdown force <span class="html-italic">F</span> as a function of drawdown velocity <span class="html-italic">v</span> in dependence of TEC content; (<b>b</b>) Relative draw resonance as a function of drawdown velocity <span class="html-italic">v</span> in dependence of TEC content (scattered area covers five measurements).</p>
Full article ">Figure 12
<p>Expansion ratio <span class="html-italic">d</span><sub>f</sub>/<span class="html-italic">d</span><sub>p</sub> of extrusion foamed CA strands using 1% Unicell TS at 40 bar as a function of TEC content.</p>
Full article ">Figure 13
<p>Influence of plasticizer content on foam morphology of CA (1% Unicell TS, <span class="html-italic">p</span><sub>die</sub> 40 bar, <span class="html-italic">T</span><sub>die</sub> 185 °C): (<b>a</b>) 15 wt % TEC; (<b>b</b>) 20 wt % TEC; and (<b>c</b>) 25 wt % TEC.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop