Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (75)

Search Parameters:
Keywords = random and block-random copolymers

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 3368 KiB  
Article
The Role of Interfacial Effects in the Impedance of Nanostructured Solid Polymer Electrolytes
by Martino Airoldi, Ullrich Steiner and Ilja Gunkel
Batteries 2024, 10(11), 401; https://doi.org/10.3390/batteries10110401 - 12 Nov 2024
Viewed by 737
Abstract
The role of interfacial effects on an ion-conducting poly(styrene-b-ethylene oxide) (PS-b-PEO or SEO) diblock copolymer doped with lithium bis(trifluoromethanesulfonyl) imide (LiTFSI) was investigated by electrochemical impedance spectroscopy (EIS). Coating the surface of commonly used stainless steel electrodes with a [...] Read more.
The role of interfacial effects on an ion-conducting poly(styrene-b-ethylene oxide) (PS-b-PEO or SEO) diblock copolymer doped with lithium bis(trifluoromethanesulfonyl) imide (LiTFSI) was investigated by electrochemical impedance spectroscopy (EIS). Coating the surface of commonly used stainless steel electrodes with a specific random copolymer brush increases the measured ionic conductivity by more than an order of magnitude compared to the uncoated electrodes. The increase in ionic conductivity is related to the interfacial structure of the block copolymer domain morphology on the electrode surface. We show that the impedance associated with the electrode–electrolyte interface can be detected using nonmetallic electrodes, allowing us to distinguish the ionic conductivity behaviors of the bulk electrolyte and the interfacial layers for both as-prepared and annealed samples. Full article
Show Figures

Figure 1

Figure 1
<p>Equivalent circuits used to fit the Nyquist plots showing single or multiple Randles circuits for the case of stainless steel (<b>a</b>) and silica (<b>b</b>) substrates.</p>
Full article ">Figure 2
<p>Nyquist plot of a PS-<span class="html-italic">b</span>-PEO SPE layer sandwiched between two pristine stainless steel electrodes (<b>a</b>) and those coated with statistical P(S-<span class="html-italic">r</span>-2VP) copolymer brushes (<b>b</b>). The red semicircles stem from single equivalent Randles circuits, which were used to fit the data (shown in the inset), yielding the ionic conductivity vs. temperature (<b>c</b>).</p>
Full article ">Figure 3
<p>Schematic BCP arrangement on preferential and neutral surfaces. (<b>a</b>) BCP layers reorganize into a substrate-parallel lamellar stack adjacent to the pristine silicon wafer substrate while exhibiting an isotropic orientation away from the substrate. (<b>b</b>) Rendering the substrate neutral by a polymer brush results in an edge-on alignment of the BCP lamellae close to the substrate. The two mixed morphologies impact near-surface impedances and ion access to the substrate. Equivalent BCP surface reconstruction on preferential (<b>c</b>) and neutral (<b>d</b>) rougher steel substrates has a less well-defined response of impedance analysis than the smooth silicon surfaces in (<b>a</b>,<b>b</b>). The brush in (<b>d</b>) is omitted for clarity.</p>
Full article ">Figure 4
<p>Nyquist plots of a PS-<span class="html-italic">b</span>-PEO SPE layer sandwiched between two pristine silicon wafer electrodes (<b>a</b>) and between two silicon wafers coated with statistical P(S-<span class="html-italic">r</span>-2VP) copolymer brushes (<b>b</b>). The two semi-circles (dashed lines) result from two Randles circuits shown in the inset. This equivalent circuit was used for data fitting to obtain the ionic conductivity values vs. temperature, which are shown together with the temperature-dependent ionic conductivities from <a href="#batteries-10-00401-f002" class="html-fig">Figure 2</a> (green and orange symbols for pristine and brush-covered surfaces). The blue symbols denote the two time constants and the corresponding conductivities of the pristine silicon wafers, while the black symbols represent the brush-covered silicon wafers (<b>c</b>).</p>
Full article ">Figure 5
<p>Bode plot of PS-<span class="html-italic">b</span>-PEO solid polymer electrolyte sandwiched between pristine (<b>a</b>) and statistical P(S-<span class="html-italic">r</span>-2VP) copolymer brush-coated (<b>b</b>) stainless steel electrodes. (<b>c</b>,<b>d</b>) show measurements for SPE layers sandwiched between pristine and brush-coated silicon wafers, respectively. The vertical orange, green, and red lines highlight the peaks in the phase signals in (<b>a</b>–<b>d</b>). The fitting profiles are plotted in red.</p>
Full article ">Figure 6
<p>Normalized imaginary component of the electric modulus (<b>a</b>) and tan<math display="inline"><semantics> <mrow> <mo>(</mo> <mi>δ</mi> <mo>)</mo> </mrow> </semantics></math> (<b>b</b>) of PS-<span class="html-italic">b</span>-PEO BCP electrolyte sandwiched between pristine (blue dotted line), brush (blue line) stainless steel electrodes, and between pristine (black dotted line), brush-coated silicon wafers (black line). The black arrow highlights the high-frequency shift upon brush functionalization of the substrate, and the green arrow indicates the presence of a new process occurring for the brush-covered silicon substrates.</p>
Full article ">Figure 7
<p>(<b>a</b>) Evolution of the PS-<span class="html-italic">b</span>-PEO solid polymer electrolyte conductivity sandwiched between two brush-covered silicon wafers before and after thermal annealing. The two different time constants are marked with empty and full symbols. The black, red, blue, and green lines correspond, respectively, to as-cast and annealed at 130 °C, 160 °C, and 190 °C. (<b>b</b>) Two ionic conductivity time constant processes (open black symbols) for the BPC SPE sandwiched between two brush-covered Si wafers, measured at 80 °C, normalized by the PEO volume fraction in the BCP and the values measured for a PEO homopolymer with comparable molecular weight, 100 kg/mol, doped with LiTFSI salt (<a href="#app1-batteries-10-00401" class="html-app">Figure S7</a>) [<a href="#B63-batteries-10-00401" class="html-bibr">63</a>], plotted as a function of thermal annealing temperature. Solid and dashed blue lines correspond to the single time constants of the same SEO copolymer sandwiched between pristine and brush-coated stainless steel electrodes, respectively.</p>
Full article ">
19 pages, 5132 KiB  
Article
Double Hydrophilic Hyperbranched Copolymer-Based Lipomer Nanoparticles: Copolymer Synthesis and Co-Assembly Studies
by Angelica Maria Gerardos and Stergios Pispas
Polymers 2024, 16(22), 3129; https://doi.org/10.3390/polym16223129 - 9 Nov 2024
Viewed by 488
Abstract
Double hydrophilic, random, hyperbranched copolymers were synthesized via reversible addition–fragmentation chain transfer (RAFT) polymerization of oligo(ethylene glycol) methyl ether methacrylate (OEGMA) and 2-(dimethylamino)ethyl methacrylate (DMAEMA) utilizing ethylene glycol dimethacrylate (EGDMA) as the branching agent. The resulting copolymers were characterized in terms of their [...] Read more.
Double hydrophilic, random, hyperbranched copolymers were synthesized via reversible addition–fragmentation chain transfer (RAFT) polymerization of oligo(ethylene glycol) methyl ether methacrylate (OEGMA) and 2-(dimethylamino)ethyl methacrylate (DMAEMA) utilizing ethylene glycol dimethacrylate (EGDMA) as the branching agent. The resulting copolymers were characterized in terms of their molecular weight and dispersity using size exclusion chromatography (SEC), and their chemical structure was confirmed using FT-IR and 1H-NMR spectroscopy techniques. The choice of the two hydrophilic blocks and the design of the macromolecular structure allowed the formation of self-assembled nanoparticles, partially due to the pH-responsive character of the DMAEMA segments and their interaction with -COOH end groups remaining from the chain transfer agent. The copolymers showed pH-responsive properties, mainly due to the protonation–deprotonation equilibria of the DMAEMA segments. Subsequently, a nanoscopic polymer–lipid (lipomer) mixed system was formulated by complexing the synthesized copolymers with cosmetic amphiphilic emulsifiers, specifically glyceryl stearate (GS) and glyceryl stearate citrate (GSC). This study aims to show that developing lipid–polymer hybrid nanoparticles can effectively address the limitations of both liposomes and polymeric nanoparticles. The effects of varying the ionic strength and pH on stimuli-sensitive polymeric and mixed polymer–lipid nanostructures were thoroughly investigated. To achieve this, the structural properties of the hybrid nanoparticles were comprehensively characterized using physicochemical techniques providing insights into their size distribution and stability. Full article
(This article belongs to the Special Issue Block Copolymers: Self-Assembly and Applications, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Cosmetic emulsifiers, glyceryl stearate (<b>left</b>) and glyceryl stearate citrate (<b>right</b>).</p>
Full article ">Figure 2
<p>Synthesis route for hyperbranched copolymers R3 and R4.</p>
Full article ">Figure 3
<p>SEC curves of the hyperbranched (<b>left</b>) and linear (<b>right</b>) OEGMA<sub>950</sub>/DMAEMA copolymers.</p>
Full article ">Figure 4
<p><sup>1</sup>H-NMR spectra of hyperbranched (<b>1</b>, red line: R3/black line: R4) and linear (<b>2</b>, red line: R9/black line: R10) copolymers.</p>
Full article ">Figure 5
<p>ATR-FTIR spectra of the neat copolymers in the solid state.</p>
Full article ">Figure 6
<p>TGA curves of neat polymers (solid lines) and the first derivative (DTG, dotted lines) of each thermogram.</p>
Full article ">Figure 7
<p>Size distributions of GSC colloid, at C = 10<sup>−4</sup> g/mL, through nanoprecipitation from THF solution.</p>
Full article ">Figure 8
<p>Size distributions from DLS analysis of lipomers R3-GSC_25 (<b>a</b>), R4-GSC_25 (<b>b</b>), R9-GSC_25 (<b>c</b>), and R10-GSC_25 (<b>d</b>). (<b>e</b>) Photo of lipomer solutions. From left to right, R3-GSC_25, R4-GSC_25, R9-GSC_25, and R10-GSC_25. A characteristic bluish tint is clearly visible for the R4-GSC_25 and R9-GSC_25 lipomers.</p>
Full article ">Figure 9
<p>Size distributions from DLS analysis of lipomers R3-GSC_50 (<b>a</b>), R4-GSC_50 (<b>b</b>), R9-GSC_50 (<b>c</b>), and R10-GSC_50 (<b>d</b>).</p>
Full article ">Figure 10
<p>Size distributions from DLS analysis of lipomers R3-GSC_70 (<b>a</b>), R4-GSC_70 (<b>b</b>), R9-GSC_70 (<b>c</b>), and R10-GSC_70 (<b>d</b>).</p>
Full article ">Figure 11
<p>Fluorescence spectra of pyrene–lipomer solutions in aqueous solution. Numbers 1 and 3 are arbitrary references to the corresponding peaks of the pyrene emission spectrum.</p>
Full article ">Figure 12
<p>Size distributions from DLS analysis of lipomers R3-GSC_25 (<b>a</b>), R4-GSC_25 (<b>b</b>), R9-GSC_25 (<b>c</b>), and R10-GSC_25 (<b>d</b>) at different pH values.</p>
Full article ">Figure 13
<p>Size distributions from DLS analysis of lipomers R3-GSC_50 (<b>a</b>), R4-GSC_50 (<b>b</b>), R9-GSC_50 (<b>c</b>), and R10-GSC_50 (<b>d</b>) at different pH values.</p>
Full article ">Figure 14
<p>Size distributions from DLS analysis of lipomers R3-GSC_70 (<b>a</b>), R4-GSC_70 (<b>b</b>), R9-GSC_70 (<b>c</b>), and R10-GSC_70 (<b>d</b>) at different pH values.</p>
Full article ">Figure 15
<p>The effect of salt on the size and mass of R3 (<b>a</b>), R4 (<b>b</b>), R9 (<b>c</b>), and R10 (<b>d</b>).</p>
Full article ">Figure 16
<p>Chemical structure of crocetin.</p>
Full article ">Figure 17
<p>UV-Vis spectra of crocus extract and crocus-loaded lipomers.</p>
Full article ">
20 pages, 7015 KiB  
Review
Recent Advances in Propylene-Based Elastomers Polymerized by Homogeneous Catalysts
by Chengkai Li, Guoqiang Fan, Gang Zheng, Rong Gao and Li Liu
Polymers 2024, 16(19), 2717; https://doi.org/10.3390/polym16192717 - 25 Sep 2024
Viewed by 1028
Abstract
Propylene-based elastomers (PBEs) have received widespread attention and research in recent years due to their structural diversity and excellent properties, and are also an important area for leading chemical companies to compete for layout, but efficient synthesis of PBEs remains challenging. In this [...] Read more.
Propylene-based elastomers (PBEs) have received widespread attention and research in recent years due to their structural diversity and excellent properties, and are also an important area for leading chemical companies to compete for layout, but efficient synthesis of PBEs remains challenging. In this paper, we review the development of PBEs and categorize them into three types, grounded in their unique chain structures, including homopolymer propylene-based elastomers (hPBEs), random copolymer propylene-based elastomers (rPBEs), and block copolymer propylene-based elastomers (bPBEs). The successful synthesis of these diverse PBEs is largely credited to the relentless innovative advancements in homogeneous catalysts (metallocene catalysts, constrained geometry catalysts, and non-metallocene catalysts). Consequently, we summarize the catalytic performance of various homogeneous catalysts employed in PBE synthesis and delve into their effect on molecular weight, molecular weight distribution, and chain structures of the resulting PBEs. In the end, based on the current academic research and industrialization status of PBEs, an outlook on potential future research directions for PBEs is provided. Full article
(This article belongs to the Section Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>Key elements for the PBEs.</p>
Full article ">Figure 2
<p>The polymerization mechanism of propylene by homogeneous catalysts (taking the polymerization of propylene as an example and some other chain transfer processes were not described for clarity).</p>
Full article ">Figure 3
<p>Catalysts for the synthesis of isotactic/atactic stereoblock PBEs.</p>
Full article ">Figure 4
<p>Schematic diagram of oscillating catalyst for the production of stereoblock PBEs.</p>
Full article ">Figure 5
<p>Plot of the propene stereoregularity ([mmmm] pentads) versus (<b>a</b>) the polymerization temperature (<span class="html-italic">T</span><sub>p</sub>) and (<b>b</b>) the monomer concentration ([<span class="html-italic">C</span><sub>3</sub>]) for catalysts <b>5</b>–<b>7</b> [<a href="#B36-polymers-16-02717" class="html-bibr">36</a>]. Copyright 1999 American Chemical Society. Stress−strain (<b>c</b>) and cyclic stress−strain (<b>d</b>) hysteresis curves of a compression-molded polypropylene specimen by catalysts <b>9</b>–<b>13</b> [<a href="#B38-polymers-16-02717" class="html-bibr">38</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 6
<p>Catalysts for the synthesis of high molecular weight <span class="html-italic"><sup>a</sup></span>PP.</p>
Full article ">Figure 7
<p>(<b>a</b>) Image of a pressed transparent disc of UHMW <span class="html-italic"><sup>a</sup></span>PP. (<b>b</b>) Cyclic stretch curves of a UHMW PP sample. The sample was extended to 120% strain cyclically. (<b>c</b>) Engineering stress–strain curves (inset: image of dog-bone samples before and after tensile testing) [<a href="#B42-polymers-16-02717" class="html-bibr">42</a>]. Copyright 2022 Elsevier B.V.</p>
Full article ">Figure 8
<p>Metallocene catalysts for the synthesis of random copolymer PBEs.</p>
Full article ">Figure 9
<p>(<b>a</b>) The effect of the initial concentration of 1-hexene on activity for the different catalytic systems. (<b>b</b>) The stress–strain curves of the copolymers [<a href="#B48-polymers-16-02717" class="html-bibr">48</a>]. Copyright 2005 Elsevier B.V.</p>
Full article ">Figure 10
<p>Non-metallocene catalysts for the synthesis of random copolymer PBEs.</p>
Full article ">Figure 11
<p>(<b>a</b>) Stress–strain curves of the copolymers with different comonomer contents. (<b>b</b>) Cycle tensile test with 10.4 mol% comonomer incorporation [<a href="#B54-polymers-16-02717" class="html-bibr">54</a>]. Copyright 2015 Elsevier B.V.</p>
Full article ">Figure 12
<p>(<b>a</b>) Plots of comonomer incorporations and contact angles as a function of comonomer loading varying from 0.75 to 1.5 mmol. (<b>b</b>) Comparison of stress-strain curves of the <span class="html-italic"><sup>i</sup></span>PP and copolymers. (<b>c</b>) Contact angle of representative (co)polymer samples [<a href="#B55-polymers-16-02717" class="html-bibr">55</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 13
<p>(<b>a</b>) The stress–strain curves of P/C20 copolymers with different 1-eicosene incorporation. (<b>b</b>) The elastic recovery of copolymers containing 20 mol% comonomers. (<b>c</b>) The preparation PBEs through copolymerization of propene and α-olefins [<a href="#B56-polymers-16-02717" class="html-bibr">56</a>]. Copyright 2020 MDPI.</p>
Full article ">Figure 14
<p>(<b>a</b>) Sequential coordination polymerization of HPD, propylene, and HPD. (<b>b</b>) Schematic illustration of a triblock TPE. (<b>c</b>) Stress-strain curves for a triblock copolymer and <span class="html-italic"><sup>a</sup></span>PP sample of similar molecular weight [<a href="#B61-polymers-16-02717" class="html-bibr">61</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 15
<p>(<b>a</b>) Synthesis of the triblock copolymer polynorbornene-<span class="html-italic">b</span>-atactic polypropene-<span class="html-italic">b</span>-polynorbornene. (<b>b</b>) The material properties of triblock copolymers [<a href="#B62-polymers-16-02717" class="html-bibr">62</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 16
<p>Non-metallocene catalysts for the synthesis of A-B-A triblock copolymer PBEs.</p>
Full article ">Figure 17
<p>Catalysts for the synthesis of multiblock copolymer PBEs.</p>
Full article ">
15 pages, 7305 KiB  
Article
Contact Hole Shrinkage: Simulation Study of Resist Flow Process and Its Application to Block Copolymers
by Sang-Kon Kim
Micromachines 2024, 15(9), 1151; https://doi.org/10.3390/mi15091151 - 13 Sep 2024
Viewed by 1054
Abstract
For vertical interconnect access (VIA) in three-dimensional (3D) structure chips, including those with high bandwidth memory (HBM), shrinking contact holes (C/Hs) using the resist flow process (RFP) represents the most promising technology for low- [...] Read more.
For vertical interconnect access (VIA) in three-dimensional (3D) structure chips, including those with high bandwidth memory (HBM), shrinking contact holes (C/Hs) using the resist flow process (RFP) represents the most promising technology for low-k1 (where CD=k1λ/NA,CD is the critical dimension, λ is wavelength, and NA is the numerical aperture). This method offers a way to reduce dimensions without additional complex process steps and is independent of optical technologies. However, most empirical models are heuristic methods and use linear regression to predict the critical dimension of the reflowed structure but do not account for intermediate shapes. In this research, the resist flow process (RFP) was modeled using the evolution method, the finite-element method, machine learning, and deep learning under various reflow conditions to imitate experimental results. Deep learning and machine learning have proven to be useful for physical optimization problems without analytical solutions, particularly for regression and classification tasks. In this application, the self-assembly of cylinder-forming block copolymers (BCPs), confined in prepatterns of the resist reflow process (RFP) to produce small contact hole (C/H) dimensions, was described using the self-consistent field theory (SCFT). This research paves the way for the shrink modeling of the enhanced resist reflow process (RFP) for random contact holes (C/Hs) and the production of smaller contact holes. Full article
(This article belongs to the Special Issue Recent Advances in Micro/Nano-Fabrication)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the resist flow process (RFP): (<b>a</b>) spin coating; (<b>b</b>) contact hole (C/H) patterns after development process; and (<b>c</b>) contact hole (C/H) patterns after thermal reflow. The structure consists of a silicon substrate and resist.</p>
Full article ">Figure 2
<p>(<b>a</b>) Scanning electron microscope (SEM) images of contact holes (C/Hs) and (<b>b</b>) a graph of critical dimension (CD) with a function of temperature and duty ratio after the thermal resist process. The upper labels of the images in <a href="#micromachines-15-01151-f002" class="html-fig">Figure 2</a>a represent the aspect ratio, which is the ratio between the contact hole size and the pitch size.</p>
Full article ">Figure 3
<p>Simulation results of Surface Evolver: (<b>a</b>) surface energy and surface area for one contact hole pattern and nine contact holes in terms of simulation times, one contact hole and nine contact holes (<b>b</b>,<b>e</b>) before the resist flow process (RFP), at surface areas of (<b>c</b>) 38.3685, (<b>d</b>) 32.9430, (<b>f</b>) 155.0468, and (<b>g</b>) 144.5697 after the resist flow process (RFP), respectively.</p>
Full article ">Figure 4
<p>Heat transfer analysis generated through MATLAB: (<b>a</b>) a block with the finite element mash displayed, (<b>b</b>) a temperature contour with constant thermal conductivity, and (<b>c</b>) a plot of simulated bake cycle profiles with constant thermal conductivity (or constant) and temperature-dependent thermal conductivity (or variable), respectively. Transfer time is 1 s.</p>
Full article ">Figure 5
<p>Static structural deformation of one contact hole (C/H) due to heat transfer, generated through ANSYS: (<b>a</b>) the structure of a <math display="inline"><semantics> <mrow> <mn>5</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> <mo>×</mo> <mn>5</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> <mo>×</mo> <mn>0.8</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> <mo> </mo> <mi mathvariant="normal">b</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">p</mi> <mi mathvariant="normal">l</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">e</mi> </mrow> </semantics></math>, a <math display="inline"><semantics> <mrow> <mn>5</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> <mo>×</mo> <mn>5</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> <mo>×</mo> <mn>5.6</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> upper plate, and an emptied cylinder with a height of <math display="inline"><semantics> <mrow> <mn>5</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> and a radius of <math display="inline"><semantics> <mrow> <mn>1.05</mn> <mo> </mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, and plots of (<b>b</b>) mashing and deformation at (<b>c</b>) 0.5 s and (<b>d</b>) 1 s simulation times.</p>
Full article ">Figure 6
<p>Phase change of a contact hole (C/H) sidewall due to gravity from solid to liquid during thermal reflow, generated using the ANSYS Fluent (Student Edition 2024): melting boundaries at (<b>a</b>) 0 steps, (<b>b</b>) 1 step, (<b>c</b>) 30 steps, (<b>d</b>) 50 steps, (<b>e</b>) 100 steps, and (<b>f</b>) 150 steps.</p>
Full article ">Figure 7
<p>Experimental and simulation results for thermal reflow biases depending on temperatures and duty ratios in a 193-nm argon fluoride (ArF) chemically amplified resist (CAR): the results of (<b>a</b>) an orthogonal fitting function, linear regressions from (<b>b</b>) a Python program and (<b>c</b>) Scikit-Learn, and (<b>d</b>) a convolutional neural network (CNN). “Exp” and “Fit” represent the experimental results and simulation results, respectively.</p>
Full article ">Figure 8
<p>Classifications from a logistic regression in deep learning and a decision tree in machine learning: (<b>a</b>,<b>b</b>) are binary and multiple classifications from TensorFlow, respectively, (<b>c</b>,<b>d</b>) are the decision surfaces of a decision tree for the binary class with the max depth of two and the multiple class with a max depth of three, respectively, and (<b>e</b>) a decision tree for multiple classes with a max depth of four.</p>
Full article ">Figure 9
<p>Multi-class classification using a support vector machine (SVM): (<b>a</b>) support vector classification (SVC) with a linear kernel, (<b>b</b>) LinearSVC (linear kernel), (<b>c</b>) support vector classification (SVC) with a radial basis function (RBF) kernel, and (<b>d</b>) support vector classification (SVC) with a polynomial (degree 3) kernel.</p>
Full article ">Figure 10
<p>Classification of resist flow process (RFP) side images using a convolutional neural network (CNN). For the upper labels of the images, the first number represents the pattern type from experimental results, and the second number represents the predicted pattern type from a convolutional neural network (CNN).</p>
Full article ">Figure 11
<p>Schematic representation of the self-assembly of cylinder-forming block copolymers (BCPs) confined in cylindrical prepatterns: (<b>a</b>) contact hole (C/H) shrinkage patterns after thermal reflow (baking), (<b>b</b>) spin coating for PS-b-PMMA block copolymers (BCPs), (<b>c</b>) thermal annealing, and (<b>d</b>) wet development.</p>
Full article ">Figure 12
<p>Comparison of the simulation results (Sim.) using rectangular guiding patterns <math display="inline"><semantics> <mrow> <mo mathvariant="normal">(</mo> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">D</mi> </mrow> <mrow> <mi mathvariant="normal">g</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">d</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">g</mi> </mrow> </msub> <mo mathvariant="normal">)</mo> </mrow> </semantics></math> to the experimental results (Exp.) using cylindrical guiding patterns <math display="inline"><semantics> <mrow> <mo mathvariant="normal">(</mo> <msub> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">D</mi> </mrow> <mrow> <mi mathvariant="normal">g</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">d</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">g</mi> </mrow> </msub> <mo mathvariant="normal">)</mo> </mrow> </semantics></math> from Reference [<a href="#B49-micromachines-15-01151" class="html-bibr">49</a>] for the resulting morphologies obtained after the directed self-assembly (DSA) of cylindrical PS-b-PMMA block copolymers (BCPs). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> is a natural period of the block copolymer (BCP), which is directly correlated to the molecular weight of the block copolymer (BCP).</p>
Full article ">
15 pages, 9036 KiB  
Review
Substrate Neutrality for Obtaining Block Copolymer Vertical Orientation
by Kaitlyn Hillery, Nayanathara Hendeniya, Shaghayegh Abtahi, Caden Chittick and Boyce Chang
Polymers 2024, 16(12), 1740; https://doi.org/10.3390/polym16121740 - 19 Jun 2024
Viewed by 838
Abstract
Nanopatterning methods utilizing block copolymer (BCP) self-assembly are attractive for semiconductor fabrication due to their molecular precision and high resolution. Grafted polymer brushes play a crucial role in providing a neutral surface conducive for the orientational control of BCPs. These brushes create a [...] Read more.
Nanopatterning methods utilizing block copolymer (BCP) self-assembly are attractive for semiconductor fabrication due to their molecular precision and high resolution. Grafted polymer brushes play a crucial role in providing a neutral surface conducive for the orientational control of BCPs. These brushes create a non-preferential substrate, allowing wetting of the distinct chemistries from each block of the BCP. This vertically aligns the BCP self-assembled lattice to create patterns that are useful for semiconductor nanofabrication. In this review, we aim to explore various methods used to tune the substrate and BCP interface toward a neutral template. This review takes a historical perspective on the polymer brush methods developed to achieve substrate neutrality. We divide the approaches into copolymer and blended homopolymer methods. Early attempts to obtain neutral substrates utilized end-grafted random copolymers that consisted of monomers from each block. This evolved into side-group-grafted chains, cross-linked mats, and block cooligomer brushes. Amidst the augmentation of the chain architecture, homopolymer blends were developed as a facile method where polymer chains with each chemistry were mixed and grafted onto the substrate. This was largely believed to be challenging due to the macrophase separation of the chemically incompatible chains. However, innovative methods such as sequential grafting and BCP compatibilizers were utilized to circumvent this problem. The advantages and challenges of each method are discussed in the context of neutrality and feasibility. Full article
(This article belongs to the Special Issue Block Copolymers: Synthesis, Self-Assembly and Application)
Show Figures

Figure 1

Figure 1
<p>Schematic of nanofabrication process of chemically patterned substrates [<a href="#B29-polymers-16-01740" class="html-bibr">29</a>] highlighting the direct assembly of a BCP process using various forms of neutrality: (<b>a</b>) homopolymer brushes; (<b>b</b>) mixed homopolymer brushes; (<b>c</b>) random copolymer brushes; (<b>d</b>) side-chain brushes; (<b>e</b>) ternary homopolymer brushes; (<b>f</b>) cross-linked polymer mats; (<b>g</b>) block cooligomer brushes. Colors represent polymers with distinct repeat units. Adapted with permission from references [<a href="#B1-polymers-16-01740" class="html-bibr">1</a>,<a href="#B29-polymers-16-01740" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Left: Simplified schematics of directed self-assembly routes. Colors represent polymers with distinct repeat units. Right: (<b>a</b>) Nitroxide-mediated radical polymerization for hydroxyl-terminated end-functionalized RCPs inspired by Mansky’s study. (<b>b</b>) Alternative confinement of a BCP thin-film using RCP derivatives [<a href="#B39-polymers-16-01740" class="html-bibr">39</a>]. Adapted with permission from reference [<a href="#B39-polymers-16-01740" class="html-bibr">39</a>].</p>
Full article ">Figure 3
<p>PS and PMMA RCPs of (<b>a</b>) end-hydroxyl functionalized brush, (<b>b</b>) side-chain hydroxyl-containing brush, (<b>c</b>) schematic of (<b>a</b>) and (<b>b</b>), respectively, based on the work of Nealey and Gopalan, followed by an SEM image of BCP of PS-<span class="html-italic">b</span>-PMMA (52-<span class="html-italic">b</span>-52) atop a side-chain brush of f<sub>St</sub> = 0.58, f<sub>MMA</sub> = 0.41, and f<sub>HEMA</sub> = 0.01 [<a href="#B40-polymers-16-01740" class="html-bibr">40</a>]. Adapted with permission from reference [<a href="#B40-polymers-16-01740" class="html-bibr">40</a>].</p>
Full article ">Figure 4
<p>Epoxy-containing RCP for use as cross-linking neutral mat with a schematic of achieved neutrality of BCP of PS-<span class="html-italic">b</span>-PMMA (52-<span class="html-italic">b</span>-52) atop the cross-linked neutral mat. f<sub>GMA</sub> = 0.01 for all images and varies by styrene fractions such that: (<b>a</b>) f<sub>St</sub> = 0.48, (<b>b</b>) f<sub>St</sub> = 0.53, (<b>c</b>) f<sub>St</sub> = 0.56, (<b>d</b>) f<sub>St</sub> = 0.59, and (<b>e</b>) f<sub>St</sub> = 0.63, scale bar represents 200 nm [<a href="#B43-polymers-16-01740" class="html-bibr">43</a>]. Adapted with permission from reference [<a href="#B43-polymers-16-01740" class="html-bibr">43</a>].</p>
Full article ">Figure 5
<p>(Left) Block cooligomer brushes (1.6–2.5 kg/mol) of f<sub>St</sub> = 0.64 grafted onto the substrate prior to annealing the BCP of PS-<span class="html-italic">b</span>-PMMA (52-<span class="html-italic">b</span>-52) atop. (Right) Nitroxide-mediated polymerization of O(S-<span class="html-italic">b</span>-M<span class="html-italic">r</span>H) block cooligomer [<a href="#B21-polymers-16-01740" class="html-bibr">21</a>]. Adapted with permission from reference [<a href="#B21-polymers-16-01740" class="html-bibr">21</a>].</p>
Full article ">Figure 6
<p>Water contact angle measurements of hydroxy-terminated PS (Mn = 3.8 kg/mol) and PMMA (Mn = 4.4 kg/mol) as a function of PS composition with corresponding optical microscopy images of 80 nm thick film with equal weights of PS = 22 kg/mol and PMMA = 23 kg/mol annealed atop PS and PMMA brushes with compositions (<b>a</b>) 100% PMMA-OH, (<b>b</b>) 60% PS, (<b>c</b>) 80% PS, and (<b>d</b>) 100% PS [<a href="#B44-polymers-16-01740" class="html-bibr">44</a>]. Adapted with permission from reference [<a href="#B44-polymers-16-01740" class="html-bibr">44</a>].</p>
Full article ">Figure 7
<p>(Left) Water contact angles of PS brushes before modification (grey), after PMMA-OH 20 (kg/mol) insertion (orange), and PS brushes treated with PMMA 20 (20 kg/mol) (teal). (Right) SEM images of BCP PS-<span class="html-italic">b</span>-PMMA (52-<span class="html-italic">b</span>-52) annealed on the inserted brushes of PMMA in PS with the relative Mn as follows: PS:PMMA (kg/mol) (<b>a</b>) 3:20, (<b>b</b>) 6:20, (<b>c</b>) 9:20, and (<b>d</b>) 20:20 [<a href="#B14-polymers-16-01740" class="html-bibr">14</a>]. Adapted with permission from reference [<a href="#B14-polymers-16-01740" class="html-bibr">14</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) SEM images of self-assembled PS-<span class="html-italic">b</span>-PMMA (52k-<span class="html-italic">b</span>-52k) on homopolymer brushes made from a blend solution (1 wt%) containing 70% BCP blender (5k-<span class="html-italic">b</span>-5k) and 30% homopolymers of PS-OH (6 k) and PMMA-OH (6 k) before rinsing. The following images display these ratios of the 30% homopolymer brushes of PS:PMMA: (<b>i</b>) 6:4, (<b>ii</b>) 5:5, and (<b>iii</b>) 4:6 [<a href="#B23-polymers-16-01740" class="html-bibr">23</a>]. (<b>b</b>) SEM images of PS-<span class="html-italic">b</span>-PMMA (50k-<span class="html-italic">b</span>-50k) on long-chain binary homopolymer-blend brushes grafted without a BCP blender using PS-OH (16 kg/mol) and PMMA-OH (15 kg/mol). The following images represent these ratios of the cast blend PS:PMMA: (<b>i</b>) 85:15, (<b>ii</b>) 80:20, and (<b>iii</b>) 75:25 [<a href="#B15-polymers-16-01740" class="html-bibr">15</a>]. Adapted with permission from references [<a href="#B15-polymers-16-01740" class="html-bibr">15</a>,<a href="#B23-polymers-16-01740" class="html-bibr">23</a>].</p>
Full article ">Figure 9
<p>(Left): Schematic of two-step insertion process of homopolymers for a neutral surface using PS-OH (17.4 kg/mol) and PDMS-OH (17.8 kg/mol) to orient cylinder-forming BCPs: PS-b-PDMS [<a href="#B45-polymers-16-01740" class="html-bibr">45</a>]. (Right): Tapping mode SPM topographical images of PS-OH (71.6% surface composition) and PDMS-OH (28.4% surface composition) with schematics depicting expected brush and amplitude fluctuations. (Top) is a weak PS-selective solvent, (middle) is a moderately PS-selective solvent, and (bottom) is a highly PS-selective solvent [<a href="#B68-polymers-16-01740" class="html-bibr">68</a>]. Adapted with permission from reference [<a href="#B68-polymers-16-01740" class="html-bibr">68</a>].</p>
Full article ">
17 pages, 37425 KiB  
Article
Melting Behaviors of Bio-Based Poly(propylene 2,5-furan dicarboxylate)-b-poly(ethylene glycol) Co Polymers Related to Their Crystal Morphology
by Ouyang Shi, Peng Li, Chao Yang, Haitian Jiang, Liyue Qin, Wentao Liu, Xiaolin Li and Zhenming Chen
Polymers 2024, 16(1), 97; https://doi.org/10.3390/polym16010097 - 28 Dec 2023
Viewed by 1237
Abstract
In this experiment, a series of poly(propylene 2,5-furan dicarboxylate)-b-poly(ethylene glycol) (PPFEG) copolymers with different ratios were synthesized using melt polycondensation of dimethylfuran-2,5-dicarboxylate (DMFD), 1,3-propanediol (PDO) and poly(ethylene glycol) (PEG). The effect of PEG content on the crystallization behavior of the poly(propylene 2,5-furan dicarboxylate) [...] Read more.
In this experiment, a series of poly(propylene 2,5-furan dicarboxylate)-b-poly(ethylene glycol) (PPFEG) copolymers with different ratios were synthesized using melt polycondensation of dimethylfuran-2,5-dicarboxylate (DMFD), 1,3-propanediol (PDO) and poly(ethylene glycol) (PEG). The effect of PEG content on the crystallization behavior of the poly(propylene 2,5-furan dicarboxylate) (PPF) copolymers was investigated. For PPF, the nucleation density of the β-crystals was higher than that of α-crystals. As Tc increases, the β crystals are suppressed more, but at Tc = 140 °C, the bulk of PPF has already been converted to α crystals, which crystallize faster at higher nucleation densities, resulting in a difference in polymer properties. For this case, we chose to add a soft segment material, PEG, which led to an early multi–melt crystallization behavior of the PPF. The addition of PEG led to a decrease in the crystallization temperature of PPF, as well as a decrease in the cold crystallization peak of PPF. From the crystalline morphology, it can be seen that the addition of PEG caused the transformation of the PPF crystalline form to occur earlier. From the crystalline morphology of PPF at 155 °C, it can be observed that the ring-banded spherical crystals of the PPF appear slowly with increasing time. With the addition of PEG, spherical crystals of the ring band appeared earlier, and even appeared first, and then disappeared slowly. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structures of PPF and PPFEG copolyesters.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR and molecular weight curves of PPF and PPFEG copolyesters.</p>
Full article ">Figure 3
<p>DSC curves of PPF and PPFEG copolymers.</p>
Full article ">Figure 4
<p>Melting curves of different <span class="html-italic">T</span><sub>c</sub> isothermal crystallization of PPF and PPFEG copolymers.</p>
Full article ">Figure 5
<p>Isothermal melting curve of PPF and PPFEG copolymers at the same <span class="html-italic">T</span><sub>c.</sub></p>
Full article ">Figure 6
<p>Secondary melting curves of PPF and PPFEG copolymers at different cooling rates peaks, depending on the corresponding peak temperature. Secondary melting of PPF and PPFEG copolymers after cooling to 0 °C at different cooling rate curves.</p>
Full article ">Figure 7
<p>POM plots of PPF at different <span class="html-italic">T</span><sub>cs</sub> and crystallization times.</p>
Full article ">Figure 8
<p>POM plots of PPFEG5 at different <span class="html-italic">T</span><sub>cs</sub> and crystallization times.</p>
Full article ">Figure 9
<p>POM plots of PPFEG10 at different <span class="html-italic">T</span><sub>cs</sub> and crystallization times.</p>
Full article ">Figure 10
<p>POM plots of PPFEG20 at different <span class="html-italic">T</span><sub>cs</sub> and crystallization times.</p>
Full article ">Figure 11
<p>POM plots of PPF at 155 °C and different crystallization times.</p>
Full article ">Figure 12
<p>POM plots of PPFEG5 at 145 °C and different crystallization times.</p>
Full article ">Figure 13
<p>POM plots of PPFEG20 at 145 °C and different crystallization times.</p>
Full article ">Figure 14
<p>WAXD diagram of PEG.</p>
Full article ">Figure 15
<p>WAXD diagram of isothermal crystallization of PPF and PPFEG copolymers.</p>
Full article ">Figure 16
<p>WAXD diagram of the non–isothermal crystallization of PPF and PPFEG copolymers.</p>
Full article ">
15 pages, 2444 KiB  
Article
Thermoresponsive Property of Poly(N,N-bis(2-methoxyethyl)acrylamide) and Its Copolymers with Water-Soluble Poly(N,N-disubstituted acrylamide) Prepared Using Hydrosilylation-Promoted Group Transfer Polymerization
by Xiangming Fu, Yanqiu Wang, Liang Xu, Atsushi Narumi, Shin-ichiro Sato, Xiaoran Yang, Xiande Shen and Toyoji Kakuchi
Polymers 2023, 15(24), 4681; https://doi.org/10.3390/polym15244681 - 12 Dec 2023
Viewed by 1243
Abstract
The group-transfer polymerization (GTP) of N,N-bis(2-methoxyethyl)acrylamide (MOEAm) initiated by Me2EtSiH in the hydrosilylation-promoted method and by silylketene acetal (SKA) in the conventional method proceeded in a controlled/living manner to provide poly(N,N-bis(2-methoxyethyl)acrylamide) (PMOEAm) and PMOEAm [...] Read more.
The group-transfer polymerization (GTP) of N,N-bis(2-methoxyethyl)acrylamide (MOEAm) initiated by Me2EtSiH in the hydrosilylation-promoted method and by silylketene acetal (SKA) in the conventional method proceeded in a controlled/living manner to provide poly(N,N-bis(2-methoxyethyl)acrylamide) (PMOEAm) and PMOEAm with the SKA residue at the α-chain end (MCIP-PMOEAm), respectively. PMOEAm-b-poly(N,N-dimethylacrylamide) (PDMAm) and PMOEAm-s-PDMAm and PMOEAm-b-poly(N,N-bis(2-ethoxyethyl)acrylamide) (PEOEAm) and PMOEAm-s-PEOEAm were synthesized by the block and random group-transfer copolymerization of MOEAm and N,N-dimethylacrylamide or N,N-bis(2-ethoxyethyl)acrylamide. The homo- and copolymer structures affected the thermoresponsive properties; the cloud point temperature (Tcp) increasing by decreasing the degree of polymerization (x). The chain-end group in PMOEAm affected the Tcp with PMOEAmx > MCIP-PMOEAmx. The Tcp of statistical copolymers was higher than that of block copolymers, with PMOEAmx-s-PDMAmy > PMOEAmx-b-PDMAmy and PMOEAmx-s-PEOEAmy > PMOEAmx-b-PEOEAmy. Full article
(This article belongs to the Section Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>SEC traces of (<b>a</b>) PMOEAm and (<b>b</b>) MCIP-PMOEAm (eluent, DMF containing lithium chloride (0.01 mol L<sup>−1</sup>); flow rate, 1.0 mL min<sup>−1</sup>).</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra of (<b>a</b>) PMOEAm<sub>25</sub> and (<b>b</b>) MCIP-PMOEAm<sub>25</sub> in CDCl<sub>3</sub>.</p>
Full article ">Figure 3
<p>MALDI-TOF MAS spectra of (<b>a</b>) PMOEAm<sub>25</sub> with a <span class="html-italic">M</span><sub>n,SEC</sub> of 4.5 kg mol<sup>−1</sup> and a Đ of 1.13 (<a href="#polymers-15-04681-t001" class="html-table">Table 1</a>, run 1) and (<b>b</b>) MCIP-PMOEAm<sub>25</sub> with a M<sub>n,SEC</sub> of 5.1 kg mol<sup>−1</sup> and a Đ of 1.11 (<a href="#app1-polymers-15-04681" class="html-app">Table S1</a>, run 1).</p>
Full article ">Figure 4
<p>(<b>a</b>) Zero-order kinetic plots and (<b>b</b>) dependence of molar mass (<span class="html-italic">M</span><sub>n</sub>) and polydispersity index (Đ) on monomer conversion (Conv.) in the B(C<sub>6</sub>F<sub>5</sub>)<sub>3</sub>-ctalyzed GTP of MOEAm with (○) Me<sub>2</sub>EtSiH and (Δ) SKA<sup>Et</sup> ([MOEAm]<sub>0</sub>/[Me<sub>2</sub>EtSiH or SKA<sup>Et</sup>]<sub>0</sub>/[B(C<sub>6</sub>F<sub>5</sub>)<sub>3</sub>]<sub>0</sub>, 100/1/0.1; [MOEAm]<sub>0</sub>, 1.0 mol L<sup>−1</sup>).</p>
Full article ">Figure 5
<p>Dependence of <span class="html-italic">T</span><sub>cp</sub> on DP<span class="html-italic"><sub>x</sub></span> for (<b>a</b>) (○) PMOEAm<span class="html-italic"><sub>x</sub></span> and (□) MCIP-PMOEAm<span class="html-italic"><sub>x</sub></span> and (<b>b</b>) (○) PMOEAm<span class="html-italic"><sub>x</sub></span>-<span class="html-italic">s</span>-PDMAm<span class="html-italic"><sub>y</sub></span>, (□) PMOEAm<span class="html-italic"><sub>x</sub></span>-<span class="html-italic">b</span>-PDMAm<span class="html-italic"><sub>y</sub></span>, (◊) PMOEAm<span class="html-italic"><sub>x</sub></span>-<span class="html-italic">s</span>-PEOEAm<span class="html-italic"><sub>y</sub></span>, and (∆) PMOEAm<span class="html-italic"><sub>x</sub></span>-<span class="html-italic">b</span>-PEOEAm<span class="html-italic"><sub>y</sub></span>.</p>
Full article ">Figure 6
<p><sup>1</sup>H NMR spectra of (<b>a</b>) PMOEAm<sub>50</sub>-<span class="html-italic">b</span>-PDMAm<sub>50</sub> measured in D<sub>2</sub>O at 30, 50, and 70 °C and (<b>b</b>) PMOEAm<sub>50</sub>-<span class="html-italic">b</span>-PEOEAm<sub>50</sub> measured in D<sub>2</sub>O at 20, 30, and 50 °C.</p>
Full article ">Scheme 1
<p>(<b>a</b>) Synthesis of PMOEAm, PMOEAm-<span class="html-italic">b</span>-PDMAm and PMOEAm-<span class="html-italic">s</span>-PDMAm, and PMOEAm-<span class="html-italic">b</span>-PEOEAm and PMOEAm-<span class="html-italic">s</span>-PEOEAm via hydrosilylation-promoted group-transfer polymerization (GTP) and (<b>b</b>) synthesis of PMOEAm functionalized with a (methoxycarbonyl)isopropyl (MCIP) group at the α-chain end (MCIP-PMOEAm) via conventional GTP using silylketene acetal (SKA) as an initiator.</p>
Full article ">
15 pages, 2370 KiB  
Article
Investigation of Crystallization, Morphology, and Mechanical Properties of Polypropylene/Polypropylene-Polyethylene Block Copolymer Blends
by Wenjun Shao, Li-Zhi Liu, Ying Wang, Yuanxia Wang, Ying Shi and Lixin Song
Polymers 2023, 15(24), 4680; https://doi.org/10.3390/polym15244680 - 12 Dec 2023
Cited by 4 | Viewed by 2039
Abstract
Polyethylene (PE)-based elastomers are the ideal choice for enhancing the compatibility of polypropylene/polyethylene (PP/PE) blends and improving the mechanical properties of PP-based materials. However, the issue of blend systems lies in the interplay between the crystallization processes. Therefore, we investigated the crystallization behavior [...] Read more.
Polyethylene (PE)-based elastomers are the ideal choice for enhancing the compatibility of polypropylene/polyethylene (PP/PE) blends and improving the mechanical properties of PP-based materials. However, the issue of blend systems lies in the interplay between the crystallization processes. Therefore, we investigated the crystallization behavior during the cooling process of a new generation of PP/PE block copolymers (PP-b-PE) and random polypropylene (PPR, a copolymer of propylene and a small amount of ethylene or an alpha-olefin) blends using in-situ X-ray diffraction/scattering and differential scanning calorimetry (DSC) techniques. We also conducted mechanical performance tests on PPR/PP-b-PE blends at room temperature and low temperature (−5 °C). The results indicate that during the cooling process, the PP phase of PP-b-PE will follow the PPR to crystallize in advance and form a eutectic mixture, thereby enhancing the compatibility of PP/PE. Moreover, the PPR/PP-b-PE blend will form stable β-(300) crystals with excellent mechanical properties. Due to the improved compatibility of PP/PE with PP-b-PE, PE crystals are dispersed within PP crystals, providing bonding that improves the toughness of PPR under the low stiffness failure conditions of PPR/PP-b-PE blends, thereby enhancing their impact performance at low and room temperatures. This research has great significance for both recycling waste plastics and enhancing the low-temperature toughness of PPR. Full article
(This article belongs to the Special Issue Advances and Applications of Block Copolymers II)
Show Figures

Figure 1

Figure 1
<p>DSC curves of cooling (<b>a</b>) and heating (<b>b</b>) of PPR, PP-b-PE, and different proportions of PPR/PP-b-PE blends.</p>
Full article ">Figure 2
<p>Linear WAXD profiles of PPR/PP-b-PE (50/50) blends cooling process at 4 °C/min.</p>
Full article ">Figure 3
<p>Linear WAXD profiles of PPR and PP-b-PE cooling to 40 °C at 4 °C/min.</p>
Full article ">Figure 4
<p>SAXS curves of PPR, PP-b-PE, and 10%, 30%, 50%, 70%, and 90% PPR/PP-b-PE blends.</p>
Full article ">Figure 5
<p>SAXS curves of PPR, PP-b-PE, and 10%, 30%, 50%, 70%, and 90% PPR/PP-b-PE blends (<b>a</b>) Linear-(I(q)-q) and (<b>b</b>) Lorentz-q<sup>2</sup>*I-(q).</p>
Full article ">Figure 5 Cont.
<p>SAXS curves of PPR, PP-b-PE, and 10%, 30%, 50%, 70%, and 90% PPR/PP-b-PE blends (<b>a</b>) Linear-(I(q)-q) and (<b>b</b>) Lorentz-q<sup>2</sup>*I-(q).</p>
Full article ">Figure 6
<p>Long period curves of PPR/PP-b-PE (50/50) blends under difficult temperatures at 4 °C/min.</p>
Full article ">Figure 7
<p>SEM images of the fracture surface of the PPR/PP-b-PE blend in specimens with different PP-b-PE content: (<b>a</b>) PPR (<b>b</b>) PPR/PP-b-PE (90/10); (<b>c</b>) PPR//PP-b-PE (90/10).</p>
Full article ">
15 pages, 4218 KiB  
Article
Synthesis and Characterization of Side-Chain Liquid-Crystalline Block Copolymers Containing Cyano-Terminated Phenyl Benzoate Moieties
by Kaito Takahashi, Daisuke Taguchi, Takashi Kajitani, Takanori Fukushima, Shoichi Kubo and Atsushi Shishido
Molecules 2023, 28(23), 7849; https://doi.org/10.3390/molecules28237849 - 29 Nov 2023
Cited by 1 | Viewed by 1297
Abstract
Block copolymers, known for their capacity to undergo microphase separation, spontaneously yield various periodic nanostructures. These precisely controlled nanostructures have attracted considerable interest due to their potential applications in microfabrication templates, conducting films, filter membranes, and other areas. However, it is crucial to [...] Read more.
Block copolymers, known for their capacity to undergo microphase separation, spontaneously yield various periodic nanostructures. These precisely controlled nanostructures have attracted considerable interest due to their potential applications in microfabrication templates, conducting films, filter membranes, and other areas. However, it is crucial to acknowledge that microphase-separated structures typically exhibit random alignment, making alignment control a pivotal factor in functional material development. To address this challenge, researchers have explored the use of block copolymers containing liquid-crystalline (LC) polymers, which offer a promising technique for alignment control. The molecular structure and LC behavior of these polymers significantly impact the morphology and alignment of microphase-separated structures. In this study, we synthesized LC diblock copolymers with cyano-terminated phenyl benzoate moieties and evaluated the microphase-separated structures and molecular alignment behaviors. The LC diblock copolymers with a narrow molecular weight distribution were synthesized by atom transfer radical polymerization. Small angle X-ray scattering measurements revealed that the block copolymers exhibit smectic LC phases and form cylinder structures with a lattice period of about 18 nm by microphase separation. The examination of block copolymer films using polarized optical microscopy and polarized UV-visible absorption spectroscopy corroborated that the LC moieties were uniaxially aligned along the alignment treatment direction. Full article
(This article belongs to the Special Issue Smart Polymeric Micro/Nanomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of the materials used in this study: (<b>a</b>) macroinitiator, (<b>b</b>) monomer M6BACP, (<b>c</b>) homopolymer PM6BACP, and (<b>d</b>) block copolymer PEO-<span class="html-italic">b</span>-PM6BACP.</p>
Full article ">Figure 2
<p>(<b>A</b>) Chemical structures representing protons used for <sup>1</sup>H NMR analysis. (<b>B</b>,<b>C</b>) Typical <sup>1</sup>H NMR spectra of PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> in CDCl<sub>3</sub> (<b>B</b>) before and (<b>C</b>) after purification. The peaks at 7.3, 6.9, and 3.7 ppm in (<b>A</b>) are derived from the residue solvent anisole. The labels a, b, b’, c, and d represent the positions of the protons (<b>A</b>) and corresponding peaks (<b>B</b>,<b>C</b>) used for the calculation of the monomer conversion and the degree of polymerization.</p>
Full article ">Figure 3
<p>SEC curves of PM6BACP (<b>1</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>2</b>), and PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>3</b>).</p>
Full article ">Figure 4
<p>DSC thermograms of PM6BACP (<b>a</b>), PEO macroinitiator (<b>b</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>c</b>), and PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>d</b>) during the third heating and cooling processes at a scan rate of 10 °C/min.</p>
Full article ">Figure 5
<p>Polarized optical microscope images during the heating process of PM6BACP (<b>a</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>b</b>), and PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>c</b>). The samples were annealed at 120 °C for 10 min and gradually cooled to room temperature prior to the observation. Scale bars, 20 µm. White crossed arrows show the direction of the polarizers.</p>
Full article ">Figure 6
<p>SAXS profiles of PM6BACP (<b>1</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>2</b>), and PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>3</b>).</p>
Full article ">Figure 7
<p>POM images of PM6BACP (<b>a</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>b</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>c</b>), and rubbed polyimide alignment layer (<b>d</b>). White crossed arrows show the direction of the polarizers. Black arrows show the rubbing direction. Scale bars, 200 µm.</p>
Full article ">Figure 8
<p>Polarized UV-vis absorption spectra of PM6BACP (<b>a</b>), PEO-<span class="html-italic">b</span>-PM6BACP<sub>46</sub> (<b>b</b>), and PEO-<span class="html-italic">b</span>-PM6BACP<sub>106</sub> (<b>c</b>). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mo>∥</mo> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mo>⊥</mo> </mrow> </msub> </mrow> </semantics></math> are absorbances parallel and perpendicular to the rubbing direction. Dashed and straight lines show the spectra before and after annealing, respectively.</p>
Full article ">Figure 9
<p>Schematic illustrations of the preparation of polymer films.</p>
Full article ">
14 pages, 13202 KiB  
Article
Effects of Block Copolymer Terminal Groups on Toughening Epoxy-Based Composites: Microstructures and Toughening Mechanisms
by Gang Li, Wenjie Wu, Xuecheng Yu, Ruoyu Zhang, Rong Sun, Liqiang Cao and Pengli Zhu
Micromachines 2023, 14(11), 2112; https://doi.org/10.3390/mi14112112 - 17 Nov 2023
Cited by 3 | Viewed by 1542
Abstract
Despite the considerable research attention paid to block copolymer (BCP)-toughened epoxy resins, the effects of their terminal groups on their phase structure are not thoroughly understood. This study fills this gap by closely examining the effects of amino and carboxyl groups on the [...] Read more.
Despite the considerable research attention paid to block copolymer (BCP)-toughened epoxy resins, the effects of their terminal groups on their phase structure are not thoroughly understood. This study fills this gap by closely examining the effects of amino and carboxyl groups on the fracture toughness of epoxy resins at different temperatures. Through the combination of scanning electron microscopy and digital image correlation (DIC), it was found that the amino-terminated BCP was capable of forming a stress-distributing network in pure epoxy resin, resulting in better toughening effects at room temperature. In a 60 wt.% silica-filled epoxy composite system, the addition of a carboxyl-terminated BCP showed little toughening effect due to the weaker filler/matrix interface caused by the random dispersion of the microphase of BCPs and distributed silica. The fracture toughness of the epoxy system at high temperatures was not affected by the terminal groups, regardless of the addition of silica. Their dynamic mechanical properties and thermal expansion coefficients are also reported in this article. Full article
(This article belongs to the Special Issue Microelectronics Assembly and Packaging: Materials and Technologies)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of (<b>a</b>) epoxy resin, (<b>b</b>) curing agent, and (<b>c</b>) BCPs.</p>
Full article ">Figure 2
<p>Schematic presentation of the processing and fabrication of composites.</p>
Full article ">Figure 3
<p>(<b>a</b>) Modulus and Tg of epoxy/BCPs; (<b>b</b>) modulus and Tg of epoxy/BCPs with 60 wt.% silica.</p>
Full article ">Figure 4
<p>TMA results of epoxy/BCPs: (<b>a</b>) 0% silica system, (<b>b</b>) 60 wt.% silica system.</p>
Full article ">Figure 5
<p>Fracture toughness of (<b>a</b>) epoxy/BCPs at 25 °C, (<b>b</b>) epoxy/BCPs at 150 °C, (<b>c</b>) epoxy/BCPs with 60 wt.% silica at 25 °C, and (<b>d</b>) epoxy/BCPs with 60 wt.% silica at 150 °C.</p>
Full article ">Figure 6
<p>Fracture surface of the SENB specimens at 25 °C: (<b>a</b>,<b>b</b>) fracture surface of epoxy; (<b>c</b>,<b>d</b>) fracture surface of EP/Amino-BCP; (<b>e</b>,<b>f</b>) fracture surface of EP/Carboxyl-BCP.</p>
Full article ">Figure 7
<p>Curing reaction and material structure description: (<b>a</b>) epoxy resin/amino-terminated BCP system and (<b>b</b>) epoxy resin/carboxyl-terminated BCP system.</p>
Full article ">Figure 8
<p>Proposed mechanism for explaining the increase in fracture toughness of the epoxy resin/carboxyl-terminated BCP system.</p>
Full article ">Figure 9
<p>Fracture surface of the SENB specimens tested at 25 and 150 °C: (<b>a</b>) epoxy, (<b>b</b>) EP/Amino-BCP, and (<b>c</b>) EP/Carboxyl-BCP were tested at 25 °C; (<b>d</b>) epoxy, (<b>e</b>) EP/Amino-BCP, and (<b>f</b>) EP/Carboxyl-BCP were tested at 150 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Fracture surface of EP/SiO<sub>2</sub> at 25 °C, (<b>b</b>) fracture surface of EP/Amino-BCP/SiO<sub>2</sub> at 25 °C, (<b>c</b>) fracture surface of epoxy/carboxyl-terminated BCP with 60 wt.% SiO<sub>2</sub> at 25 °C, (<b>d</b>) fracture surface of EP/SiO<sub>2</sub> at 150 °C, (<b>e</b>) fracture surface of EP/Amino-BCP/SiO<sub>2</sub> at 150 °C, and (<b>f</b>) fracture surface of EP/Carboxyl-BCP/SiO<sub>2</sub> at 150 °C.</p>
Full article ">Figure 11
<p>Proposed mechanism for explaining the increase in fracture toughness of EP/Carboxyl-BCP/SiO<sub>2</sub>.</p>
Full article ">Figure 12
<p>Results of crack-tip strain of (<b>a</b>) epoxy and epoxy/BCP system at maximum load, (<b>b</b>) epoxy and epoxy/BCP with 60 wt.% SiO<sub>2</sub> system at maximum load.</p>
Full article ">
12 pages, 2645 KiB  
Article
Density Fluctuations Inside an Individual Polymer Coil
by Anatoly E. Chalykh, Uliana V. Nikulova, Vladimir K. Gerasimov and Vladimir V. Matveev
Polymers 2023, 15(19), 4018; https://doi.org/10.3390/polym15194018 - 7 Oct 2023
Viewed by 918
Abstract
More than five hundred images of individual macromolecules of random styrene-butadiene copolymers and styrene-isoprene block copolymers dissolved in a polystyrene matrix were analyzed. The presence of density fluctuations inside the macromolecular coil has been established. Within the framework of the model of harmonic [...] Read more.
More than five hundred images of individual macromolecules of random styrene-butadiene copolymers and styrene-isoprene block copolymers dissolved in a polystyrene matrix were analyzed. The presence of density fluctuations inside the macromolecular coil has been established. Within the framework of the model of harmonic oscillations, the radial distribution of such density fluctuations is estimated. Full article
(This article belongs to the Section Polymer Physics and Theory)
Show Figures

Figure 1

Figure 1
<p>An electron microphotograph of an individual SBR-45 macromolecule in a polystyrene matrix (<b>a</b>) and the distribution of gray levels corresponding to it over the image area of the macromolecule (<b>b</b>), the blackening radial function (<b>c</b>), and the segment density radial distribution function (<b>d</b>).</p>
Full article ">Figure 2
<p>Average radial distribution functions of segment density (lines 1) for SBR-96 (<b>a</b>), SBR-45 (<b>b</b>), SIS-4114 (<b>c</b>), and SIS-4215 (<b>d</b>). Lines 2 correspond to calculations within the framework of Equation (1). The current radii are normalized to the radius of gyration.</p>
Full article ">Figure 3
<p>Radial distribution functions of segments for macromolecules SBR-96 (M=106.4 kDa) (<b>a</b>), SBR-45 (M = 57.6 kDa) (<b>b</b>), SIS-4114 (M = 100.9 kDa) (<b>c</b>), and SIS-4215 (M = 24.3 kDa) (<b>d</b>); 1—obtained by processing the image of a macromolecule; 2—<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">ρ</mi> </mrow> <mrow> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">q</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi mathvariant="normal">r</mi> </mrow> </mfenced> </mrow> </semantics></math> according to Equation (1) with the constants (N, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="normal">R</mi> </mrow> <mrow> <mi mathvariant="normal">g</mi> </mrow> </msub> </mrow> </semantics></math>) corresponding to each individual macromolecule. The indicated molecular weight (M) corresponds to the specific macromolecule chosen.</p>
Full article ">Figure 3 Cont.
<p>Radial distribution functions of segments for macromolecules SBR-96 (M=106.4 kDa) (<b>a</b>), SBR-45 (M = 57.6 kDa) (<b>b</b>), SIS-4114 (M = 100.9 kDa) (<b>c</b>), and SIS-4215 (M = 24.3 kDa) (<b>d</b>); 1—obtained by processing the image of a macromolecule; 2—<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">ρ</mi> </mrow> <mrow> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">q</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi mathvariant="normal">r</mi> </mrow> </mfenced> </mrow> </semantics></math> according to Equation (1) with the constants (N, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="normal">R</mi> </mrow> <mrow> <mi mathvariant="normal">g</mi> </mrow> </msub> </mrow> </semantics></math>) corresponding to each individual macromolecule. The indicated molecular weight (M) corresponds to the specific macromolecule chosen.</p>
Full article ">Figure 4
<p>Fluctuation part of the radial density distribution functions for macromolecules shown in <a href="#polymers-15-04018-f003" class="html-fig">Figure 3</a>. Explanations in the text.</p>
Full article ">Figure 5
<p>The position of the extrema of the functions shown in <a href="#polymers-15-04018-f004" class="html-fig">Figure 4</a>. The linear dependence is described using the equation <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">ω</mi> <mi mathvariant="normal">r</mi> <mo>+</mo> <mi mathvariant="sans-serif">α</mi> </mrow> </semantics></math>. Explanations in the text.</p>
Full article ">Figure 6
<p>Radial segment distribution functions for copolymers whose segment density radial distribution functions are shown in <a href="#polymers-15-04018-f003" class="html-fig">Figure 3</a> (1); 2—description by the model of harmonic oscillations (7).</p>
Full article ">Figure 6 Cont.
<p>Radial segment distribution functions for copolymers whose segment density radial distribution functions are shown in <a href="#polymers-15-04018-f003" class="html-fig">Figure 3</a> (1); 2—description by the model of harmonic oscillations (7).</p>
Full article ">Figure 7
<p>Correlation between Flory–Huggins parameter χ (<b>a</b>,<b>b</b>), number of segments N (<b>c</b>,<b>d</b>) and amplitude А for all studied individual macromolecules; 1—SBR-96, 2—SBR-45, 3—SIS-4114, and 4—SIS-4215.</p>
Full article ">Figure 8
<p>Histogram of the distribution of the amplitude of harmonic oscillations describing density fluctuations of all studied macromolecules. The continuous curve is a function of the normal logarithmic distribution.</p>
Full article ">Figure 9
<p>Correlation between the Flory–Huggins parameter χ (<b>a</b>,<b>b</b>), the number of segments N (<b>c</b>,<b>d</b>), and the harmonic frequency <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">ω</mi> </mrow> </semantics></math> for all studied individual macromolecules; 1—SBR-96, 2—SBR-45, 3—SIS-4114, and 4—SIS-4215.</p>
Full article ">Figure 10
<p>Histogram of the distribution of the frequency of harmonic oscillations describing density fluctuations for all studied macromolecules. Continuous curve is a normal distribution function.</p>
Full article ">Figure 11
<p>Histogram of the phase shift distribution from the center of mass.</p>
Full article ">
15 pages, 4031 KiB  
Article
Semi-Spontaneous Post-Crosslinking Triblock Copolymer Electrolyte for Solid-State Lithium Battery
by Zhenan Zheng, Jie Huang, Xiang Gao and Yingwu Luo
Batteries 2023, 9(9), 465; https://doi.org/10.3390/batteries9090465 - 13 Sep 2023
Cited by 1 | Viewed by 1680
Abstract
The solid polymer electrolyte is a promising candidate for solid-state lithium battery because of favorable interfacial contact, good processability and economic availability. However, its application is limited because of low ionic conductivity and insufficient mechanical strength. In this study, the delicate molecular structural [...] Read more.
The solid polymer electrolyte is a promising candidate for solid-state lithium battery because of favorable interfacial contact, good processability and economic availability. However, its application is limited because of low ionic conductivity and insufficient mechanical strength. In this study, the delicate molecular structural design was realized via controlled / “living” radical polymerization in order to decouple the trade-off between ionic conductivity and mechanical strength. The random and triblock copolymer electrolytes were designed and synthesized to investigate the influence of molecular structure on ionic conduction, while a chemical cross-linking network was constructed via a semi-spontaneous post-crosslinking reaction. Compared with a random counterpart, the triblock copolymer electrolyte presented stronger chain segment motion and a liquid-like mechanical response due to the independent ion-conducting block, resulting in significantly improved ionic conductivity (from 6.29 ± 1.11 × 10−5 to 9.57 ± 2.82 × 10−5 S cm−1 at 60 °C) and cell performance. When assembled with LiFePO4 and lithium metal electrodes, the cell with triblock copolymer electrolyte showed significantly improved rate performance (150 mAh g−1 at 1 C) and cycling life (200 cycles with 92.8% capacity retention at 1 C). This study demonstrates the advantages of molecular structure regulation on ionic conduction and mechanical support, which may provide new insights for the future design of solid polymer electrolytes. Full article
Show Figures

Figure 1

Figure 1
<p>The relationship between ionic conductivity (at 60 °C) and molar ratio of TMSiPA.</p>
Full article ">Figure 2
<p>(<b>a</b>) Illustration of the molecular structure of TRISPE and RANSPE. (<b>b</b>) DSC curves and corresponding derivative curves of TRI-SPE and RAN-SPE.</p>
Full article ">Figure 3
<p>Arrhenius plots for SPEs with different molecular structures.</p>
Full article ">Figure 4
<p>Thermogravimetry analysis curves of PEGMA homopolymer, TMSiPA homopolymer and synthesized copolymer electrolyte.</p>
Full article ">Figure 5
<p>Relationship between (<b>a</b>) complex modulus (G*), (<b>b</b>) loss angle tangent (tan δ) and test frequency of RAN-SPE and TRI-SPE.</p>
Full article ">Figure 6
<p>Electrochemical stability window of synthesized copolymer electrolyte. The test temperature was 60 °C.</p>
Full article ">Figure 7
<p>Rate performance of LiFePO<sub>4</sub>//SPE//Li half-cells fabricated by synthesized copolymer electrolytes with different molecular structures. The test temperature was 60 °C.</p>
Full article ">Figure 8
<p>Nyquist plots of LiFePO<sub>4</sub>//SPE//Li half-cells fabricated by different synthesized copolymer electrolytes after rate performance tests and at discharge state (inset zoom-in the range of 50–550 Ω). The test temperature was 60 °C.</p>
Full article ">Figure 9
<p>Cycle performances and corresponding Coulombic efficiencies of LiFePO<sub>4</sub>//SPE//Li half-cells fabricated by synthesized copolymer electrolytes. The test temperature was 60 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Cycle performance and corresponding Coulombic efficiencies of LiFePO<sub>4</sub>//SPE//Li half-cell fabricated by TRI-SPE. (<b>b</b>) Charge and discharge voltage profiles at different cycles. (<b>c</b>) Nyquist plots at different cycles and at discharge state (inset zoom-in the range of 0–700 Ω).</p>
Full article ">Scheme 1
<p>Hydrolysis and dehydration crosslinking mechanism of synthesized copolymer electrolyte.</p>
Full article ">
22 pages, 18550 KiB  
Article
Molecular Dynamics Simulation of Hydrogels Based on Phosphorylcholine-Containing Copolymers for Soft Contact Lens Applications
by Katarzyna Filipecka-Szymczyk, Malgorzata Makowska-Janusik and Wojciech Marczak
Molecules 2023, 28(18), 6562; https://doi.org/10.3390/molecules28186562 - 11 Sep 2023
Cited by 1 | Viewed by 2218
Abstract
The structure and dynamics of copolymers of 2-hydroxyethyl methacrylate (HEMA) with 2-methacryloyloxyethyl phosphorylcholine (MPC) were studied by molecular dynamics simulations. In total, 20 systems were analyzed. They differed in numerical fractions of the MPC in the copolymer chain, equal to 0.26 and 0.74, [...] Read more.
The structure and dynamics of copolymers of 2-hydroxyethyl methacrylate (HEMA) with 2-methacryloyloxyethyl phosphorylcholine (MPC) were studied by molecular dynamics simulations. In total, 20 systems were analyzed. They differed in numerical fractions of the MPC in the copolymer chain, equal to 0.26 and 0.74, in the sequence of mers, block and random, and the water content, from 0 to 60% by mass. HEMA side chains proved relatively rigid and stable in all considered configurations. MPC side chains, in contrast, were mobile and flexible. Water substantially influenced their dynamics. The copolymer swelling caused by water resulted in diffusion channels, pronounced in highly hydrated systems. Water in the hydrates existed in two states: those that bond to the polymer chain and the free one; the latter was similar to bulk water but with a lower self-diffusion coefficient. The results proved that molecular dynamics simulations could facilitate the preliminary selection of the polymer materials for specific purposes before their synthesis. Full article
(This article belongs to the Section Computational and Theoretical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A fragment of the P(MPC–<span class="html-italic">co</span>–HEMA) copolymer chain. The atoms are tagged to simplify referring.</p>
Full article ">Figure 2
<p>Radial distribution function <span class="html-italic">g</span>(<span class="html-italic">r</span>) of the C<sub>H</sub>–O<sub>H3</sub> pair of HEMA in the B13, B37, R13, and R37 copolymers. Blue line—non-hydrated polymer; black, red, green, and magenta lines—hydrated polymers containing 10, 20, 40, and 60% of water by mass, respectively.</p>
Full article ">Figure 3
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) of the C<sub>M</sub>–N<sub>M</sub> pair of MPC in the B13, B37, R13, and R37 copolymers. Blue line—non-hydrated polymer; black, red, green, and magenta lines—hydrated polymers containing 10, 20, 40, and 60% of water by mass, respectively.</p>
Full article ">Figure 4
<p>Snapshots of the simulated adjacent MPC side chains in non-hydrated block copolymer (<b>a</b>) and that containing 60% of water by mass (<b>b</b>). Hydrogen atoms and water molecules were omitted for the picture clarity. Atoms: C—grey, O—red, N—blue, and P—magenta.</p>
Full article ">Figure 5
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water oxygen (O<sub>W</sub>) and C<sub>H</sub>, O<sub>H3</sub>, and O<sub>H2</sub> of HEMA for the R37 copolymer with various water content. Black, red, green, and magenta lines—hydrated polymers containing 10, 20, 40, and 60% of water by mass, respectively.</p>
Full article ">Figure 6
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water oxygen (O<sub>w</sub>) and C<sub>M</sub>, N<sub>M</sub>, and P<sub>M</sub> of MPC for the R37 copolymer with various water content. Black, red, green, and magenta lines—hydrated polymers containing 10, 20, 40, and 60% of water by mass, respectively.</p>
Full article ">Figure 7
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water hydrogen and HEMA oxygen for the R37 copolymer with various water content. Green lines: O<sub>H1</sub>–H<sub>W</sub>, red: O<sub>H2</sub>–H<sub>W</sub>, and black: O<sub>H3</sub>–H<sub>W</sub>.</p>
Full article ">Figure 8
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water oxygen and HEMA oxygen for the R37 copolymer with various water content. Green lines: O<sub>H1</sub>–O<sub>W</sub>, red: O<sub>H2</sub>–O<sub>W</sub>, and black: O<sub>H3</sub>–O<sub>W</sub>.</p>
Full article ">Figure 9
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water hydrogen and MPC oxygen for the R37 copolymer with various water content. Black lines: O<sub>M1</sub>–H<sub>W</sub>, red: O<sub>M2</sub>–H<sub>W</sub>; green: O<sub>M3</sub>–H<sub>W</sub>, blue: O<sub>M(4,5)</sub>–H<sub>W</sub>, cyan: O<sub>M6</sub>–H<sub>W</sub>.</p>
Full article ">Figure 10
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water hydrogen and MPC nitrogen or phosphorus for the R37 copolymer with various water content. Black lines: N<sub>M</sub>–H<sub>W</sub>, red: P<sub>M</sub>–H<sub>W</sub>.</p>
Full article ">Figure 11
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water oxygen and MPC oxygen for the R37 copolymer with various water content. Black lines: O<sub>M1</sub>–O<sub>W</sub>, red: O<sub>M2</sub>–O<sub>W</sub>; green: O<sub>M3</sub>–O<sub>W</sub>, blue: O<sub>M(4,5)</sub>–O<sub>W</sub>, cyan: O<sub>M6</sub>–O<sub>W</sub>.</p>
Full article ">Figure 12
<p>Radial distribution functions <span class="html-italic">g</span>(<span class="html-italic">r</span>) for atom pairs: water oxygen and MPC nitrogen or phosphorus for the R37 copolymer with various water content. Black lines: N<sub>M</sub>–O<sub>W</sub>, red: P<sub>M</sub>–O<sub>W</sub>.</p>
Full article ">Figure 13
<p>Morphology of free volumes in R37, dry (<b>a</b>) and with different water content: 10% (<b>b</b>), 20% (<b>c</b>), 40% (<b>d</b>), and 60% (<b>e</b>) by mass for the probe of infinitesimal radius. The walls of the empty channels making up the free volumes are marked in blue.</p>
Full article ">Figure 14
<p>The fractional free volume (FFV) of non-hydrated P(MPC–<span class="html-italic">co</span>–HEMA) as determined with probes of various radii.</p>
Full article ">Figure 15
<p>The volume in the simulated R37 material available for or filled with water: dry (<b>a</b>) and hydrated copolymers containing 10% (<b>b</b>), 20% (<b>c</b>), 40% (<b>d</b>), and 60% (<b>e</b>) of water by mass. The walls of polymer channels available for the probe with a radius of 1.4 Å (i.e., a “water molecule”) are marked in blue.</p>
Full article ">Figure 16
<p>The exponents in the power law (Equation (1)) that evidence subdiffusion of the O<sub>H3</sub> (open symbols) and N<sub>M</sub> atoms (filled symbols) of R37 vs. the mass concentration of water. Lines are guides for the eye only.</p>
Full article ">Figure 17
<p>Van Hove self-correlation functions <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>G</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mover accent="true"> <mrow> <mi>r</mi> </mrow> <mo stretchy="false">→</mo> </mover> <mo>,</mo> <mi>t</mi> </mrow> </mfenced> </mrow> </semantics></math> calculated for water molecules in R37.</p>
Full article ">Figure 18
<p>“Rolling” of a spherical probe with radius <span class="html-italic">r</span> over the surface of a molecule built of two atoms with van der Waals radii <span class="html-italic">r</span><sub>1</sub> and <span class="html-italic">r</span><sub>2</sub>.</p>
Full article ">
12 pages, 6919 KiB  
Article
Morphological Evolution of Hybrid Block Copolymer Particles: Toward Magnetic Responsive Particles
by Jaeman J. Shin
Polymers 2023, 15(18), 3689; https://doi.org/10.3390/polym15183689 - 7 Sep 2023
Viewed by 1267
Abstract
The co-assembly of block copolymers (BCPs) and inorganic nanoparticles (NPs) under emulsion confinement allows facile access to hybrid polymeric colloids with controlled hierarchical structures. Here, the effect of inorganic NPs on the structure of the hybrid BCP particles and the local distribution of [...] Read more.
The co-assembly of block copolymers (BCPs) and inorganic nanoparticles (NPs) under emulsion confinement allows facile access to hybrid polymeric colloids with controlled hierarchical structures. Here, the effect of inorganic NPs on the structure of the hybrid BCP particles and the local distribution of NPs are studied, with a particular focus on comparing Au and Fe3O4 NPs. To focus on the effect of the NP core, Au and Fe3O4 NPs stabilized with oleyl ligands were synthesized, having a comparable diameter and grafting density. The confined co-assembly of symmetric polystyrene-b-poly(1,4-butadiene) (PS-b-PB) BCPs and NPs in evaporative emulsions resulted in particles with various morphologies including striped ellipsoids, onion-like particles, and their intermediates. The major difference in PS-b-PB/Au and PS-b-PB/Fe3O4 particles was found in the distribution of NPs inside the particles that affected the overall particle morphology. Au NPs were selectively localized inside PB domains with random distributions regardless of the particle morphology. Above the critical volume fraction, however, Au NPs induced the morphological transition of onion-like particles into ellipsoids by acting as an NP surfactant. For PS-b-PB/Fe3O4 ellipsoids, Fe3O4 NPs clustered and segregated to the particle/surrounding interface of the ellipsoids even at a low volume fraction, while Fe3O4 NPs were selectively localized in the middle of PB domains in a string-like pattern for PS-b-PB/Fe3O4 onion-like particles. Full article
(This article belongs to the Section Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>Illustration showing co-assembly of inorganic nanoparticles (Au NP and Fe<sub>3</sub>O<sub>4</sub> NP) with PS-<span class="html-italic">b</span>-PB BCPs by controlled solvent evaporation from toluene-in-water emulsion.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) TEM image and (<b>c</b>,<b>d</b>) size distribution histogram of (<b>a</b>,<b>c</b>) Au NPs and (<b>b</b>,<b>d</b>) Fe<sub>3</sub>O<sub>4</sub> NPs.</p>
Full article ">Figure 3
<p>TEM images showing hybrid PS-<span class="html-italic">b</span>-PB particles. Striped ellipsoids (<b>a</b>,<b>b</b>) and onion-like particles (<b>c</b>,<b>d</b>) containing (<b>a</b>,<b>c</b>) Au NP or (<b>b</b>,<b>d</b>) magnetic NPs. Fe<sub>3</sub>O<sub>4</sub> NPs were segregated at the edge of the ellipsoidal particles as indicated by red dashed circles in <a href="#polymers-15-03689-f003" class="html-fig">Figure 3</a>b. The loading amount of NPs was fixed to 7.0 vol% for both Au and Fe<sub>3</sub>O<sub>4</sub>. Particles were observed without staining. Scale bars in inset figures are 50 nm.</p>
Full article ">Figure 4
<p>TEM images showing the distribution of Au NPs in PS-<span class="html-italic">b</span>-PB/Au ellipsoids as a function of the volume fraction of NP relative to BCP (φ). (<b>a</b>) 1.4 vol%, (<b>b</b>) 3.5 vol%, and (<b>c</b>) 7.0 vol%. Particles were observed without staining. A<sub>emul/air</sub> = 26.4 cm<sup>2</sup>.</p>
Full article ">Figure 5
<p>TEM images of hybrid PS-<span class="html-italic">b</span>-PB/Au particles as a function of φ. (<b>a</b>) Pristine PS-<span class="html-italic">b</span>-PB particles, (<b>b</b>) φ = 1.4 vol%, (<b>c</b>) φ = 3.5 vol%, and (<b>d</b>) φ = 7.0 vol%. Each color of the bar chart shows the percentage of each morphology. Black: ellipsoids, red: intermediates, blue: onions. PB domains were stained with OsO<sub>4</sub>. A<sub>emul/air</sub> = 0.1 cm<sup>2</sup>.</p>
Full article ">Figure 6
<p>TEM images showing the distribution of Fe<sub>3</sub>O<sub>4</sub> NPs in PS-<span class="html-italic">b</span>-PB striped ellipsoids as a function of φ. (<b>a</b>) 1.4 vol%, (<b>b</b>) 3.5 vol%, and (<b>c</b>) 7.0 vol%. Particles were observed without staining.</p>
Full article ">Figure 7
<p>Schematic illustration showing the formation mechanism of PS-<span class="html-italic">b</span>-PB/Fe<sub>3</sub>O<sub>4</sub> ellipsoids and onion-like particles.</p>
Full article ">
17 pages, 5236 KiB  
Article
Effect of Octene Block Copolymer (OBC) and High-Density Polyethylene (HDPE) on Crystalline Morphology, Structure and Mechanical Properties of Octene Random Copolymer
by Yuan-Xia Wang, Cun-Ying Zou, Nan Bai, Qun-Feng Su, Li-Xin Song and Xian-Liang Li
Polymers 2023, 15(18), 3655; https://doi.org/10.3390/polym15183655 - 5 Sep 2023
Viewed by 1532
Abstract
Blending octene random copolymer (ORC) with other polymers is a promising approach to improving ORC mechanical properties, such as tensile strength and elongation. In this study, octene block copolymer (OBC) with lower density than ORC and high-density polyethylene (HDPE) were used to blend [...] Read more.
Blending octene random copolymer (ORC) with other polymers is a promising approach to improving ORC mechanical properties, such as tensile strength and elongation. In this study, octene block copolymer (OBC) with lower density than ORC and high-density polyethylene (HDPE) were used to blend with ORC. The effect of both OBC and HDPE on ORC was analyzed using scanning electron microscopy (SEM), differential scanning calorimetry (DSC), dynamic mechanical analysis (DMA) and small-angle X-ray scattering (SAXS). For ORC/OBC blends, a small amount of OBC can improve the crystallization ability of ORC. Meanwhile, for ORC/HDPE blends, the crystallization ability of ORC was significantly suppressed, attributed to good compatibility between ORC and HDPE as indicated by the homogeneous morphology and the disappearance of the α transition peak of ORC in ORC/HDPE blends. Therefore, the tensile strength and elongation of ORC/HDPE blends are significantly higher than those of ORC/OBC blends. For ORC/OBC/HDPE ternary blends, we found that when ORC:OBC:HDPE are at a ratio of 70:15:15, cocrystallization is achieved. Although HDPE improves the compatibility of ORC and OBC, the three-phase structure of the ternary blends can be observed through SAXS when HDPE and OBC exceed 30 wt%. Blending HDPE and OBC (≤30 wt%) could improve the mechanical property of ORC. Full article
(This article belongs to the Special Issue Characterization and Application of Block Copolymers)
Show Figures

Figure 1

Figure 1
<p>DSC cooling (10 °C/min) thermograms (<b>a</b>,<b>b</b>) of pure ORC, pure OBC, pure HDPE, ORC/OBC blends and ORC/HDPE blends.</p>
Full article ">Figure 2
<p>DSC the second heating (10 °C/min) thermograms (<b>a</b>,<b>b</b>) of pure ORC, pure OBC, pure HDPE, ORC/OBC blends and ORC/HDPE blends.</p>
Full article ">Figure 3
<p>DSC cooling (10 °C/min) thermograms and the subsequent melting (10 °C/min) thermograms of pure ORC, pure OBC, pure HDPE, ORC/OBC blends, ORC/HDPE blends and ORC/OBC/HDPE blends: (<b>a</b>–<b>c</b>) cooling thermograms; (<b>a’</b>–<b>c’</b>) the subsequent melting thermograms.</p>
Full article ">Figure 4
<p>The SEM images of cryo-fracture of pure samples: (<b>a</b>) pure ORC, (<b>b</b>) pure OBC, (<b>c</b>) pure HDPE.</p>
Full article ">Figure 5
<p>The SEM images of cryo-fracture of the blends: (<b>a</b>) ORC/OBC10; (<b>b</b>) ORC/OBC30; (<b>c</b>) ORC/OBC50; (<b>d</b>) ORC/HDPE10; (<b>e</b>) ORC/HDPE30; (<b>f</b>) ORC/HDPE50; (<b>g</b>) ORC/OBC5/HDPE5; (<b>h</b>) ORC/OBC15/HDPE15; (<b>i</b>) ORC/OBC25/HDPE25.</p>
Full article ">Figure 6
<p>Linear and Lorentz-corrected SAXS profiles of pure ORC, pure OBC, pure HDPE, ORC/OBC blends and ORC/HDPE blends. Liner SAXS profiles: (<b>a</b>) ORC/OBC, (<b>b</b>) ORC/HDPE; Lorentz-corrected SAXS profiles: (<b>a’</b>) ORC/OBC, (<b>b’</b>) ORC/HDPE.</p>
Full article ">Figure 7
<p>Linear and Lorentz-corrected SAXS profiles of theblends under the different conditions: (<b>a</b>–<b>c</b>) liner SAXS profiles of pure ORC, pure OBC, pure HDPE, ORC/OBC blends, ORC/HDPE blends, ORC/OBC/HDPE blends; (<b>a’</b>–<b>c’</b>) Lorentz-corrected SAXS profiles of pure ORC, pure OBC, pure HDPE, ORC/OBC blends, ORC/HDPE blends, ORC/OBC/HDPE blends.</p>
Full article ">Figure 8
<p>Storage modulus (<b>a</b>,<b>b</b>) and tan δ (<b>a’</b>,<b>b’</b>) of pure ORC, pure OBC, pure HDPE, ORC/OBC blends and ORC/HDPE blends.</p>
Full article ">Figure 9
<p>Storage modulus (<b>a</b>–<b>c</b>) and tan δ (<b>a’</b>–<b>c’</b>) of pure ORC, pure OBC, pure HDPE, ORC/OBC blends, ORC/HDPE blends and ORC/OBC/HDPE blends under the different conditions.</p>
Full article ">Figure 9 Cont.
<p>Storage modulus (<b>a</b>–<b>c</b>) and tan δ (<b>a’</b>–<b>c’</b>) of pure ORC, pure OBC, pure HDPE, ORC/OBC blends, ORC/HDPE blends and ORC/OBC/HDPE blends under the different conditions.</p>
Full article ">Figure 10
<p>The stress–strain curves (up to break) of (<b>a</b>) pure ORC, pure OBC, ORC/OBC blends; (<b>b</b>) pure ORC and ORC/HDPE blends; (<b>c</b>) ORC/OBC blends and ORC/OBC/HDPE blends.</p>
Full article ">
Back to TopTop