Journal articles on the topic 'Polymer brush architecture'

To see the other types of publications on this topic, follow the link: Polymer brush architecture.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Polymer brush architecture.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Choi, Jihoon, Michael J. A. Hore, Nigel Clarke, Karen I. Winey, and Russell J. Composto. "Nanoparticle Brush Architecture Controls Polymer Diffusion in Nanocomposites." Macromolecules 47, no. 7 (March 19, 2014): 2404–10. http://dx.doi.org/10.1021/ma500235v.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Gunkel, Gesine, Marie Weinhart, Tobias Becherer, Rainer Haag, and Wilhelm T. S. Huck. "Effect of Polymer Brush Architecture on Antibiofouling Properties." Biomacromolecules 12, no. 11 (November 14, 2011): 4169–72. http://dx.doi.org/10.1021/bm200943m.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Yoshikawa, Chiaki, Keita Sakakibara, Punnida Nonsuwan, Tomohiko Yamazaki, and Yoshinobu Tsujii. "Nonbiofouling Coatings Using Bottlebrushes with Concentrated Polymer Brush Architecture." Biomacromolecules 22, no. 6 (May 3, 2021): 2505–14. http://dx.doi.org/10.1021/acs.biomac.1c00247.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Aguilar-Castillo, Bethsy Adriana, Jose Luis Santos, Hanying Luo, Yanet E. Aguirre-Chagala, Teresa Palacios-Hernández, and Margarita Herrera-Alonso. "Nanoparticle stability in biologically relevant media: influence of polymer architecture." Soft Matter 11, no. 37 (2015): 7296–307. http://dx.doi.org/10.1039/c5sm01455g.

Full text
Abstract:
We contrast the behavior of nanoparticles formed by the self-assembly of polymers based on poly(ethylene glycol) (PEG) and poly(d,l-lactide), with linear, linear-dendritic and bottle-brush architectures in biologically relevant media.
APA, Harvard, Vancouver, ISO, and other styles
5

Schmitt, Michael, Chin Ming Hui, Zachary Urbach, Jiajun Yan, Krzysztof Matyjaszewski, and Michael R. Bockstaller. "Tailoring structure formation and mechanical properties of particle brush solids via homopolymer addition." Faraday Discussions 186 (2016): 17–30. http://dx.doi.org/10.1039/c5fd00121h.

Full text
Abstract:
Recent progress in the area of surface-initiated controlled radical polymerization (SI-CRP) has enabled the synthesis of polymer-grafted colloids with precise control over the architecture of grafted chains. The resulting ‘particle brush materials’ are of interest both from a fundamental as well as applied perspective because structural frustrations (associated with the tethering of chains to a curved surface) imply a sensitive dependence of the interactions between brush particles on the architecture of surface-tethered chains that offers new opportunities to design hybrid materials with novel functionalities. An important prerequisite for establishing structure–property relations in particle brush materials is to understand the role of homopolymer impurities that form, for example, by thermal self-initiation. This contribution presents a detailed discussion of the role of homopolymer additives on the structure and mechanical properties of particle brush materials. The results suggest that the dissolution of homopolymer fillers follows a two-step mechanism comprised of the initial segregation of homopolymer to the interstitial regions within the array and the subsequent swelling of the particle brush (depending on the respective degree of polymerization of brush and linear chains). Addition of even small amounts of homopolymer is found to significantly increase the fracture toughness of particle brush assembly structures. The increased resistance to failure could enable the synthesis of robust colloidal crystal type materials that can be processed into complex shapes using ‘classical’ polymer forming techniques such as molding or extrusion.
APA, Harvard, Vancouver, ISO, and other styles
6

Alves, Patrícia, Luciana Calheiros Gomes, Cesar Rodríguez-Emmenegger, and Filipe José Mergulhão. "Efficacy of A Poly(MeOEGMA) Brush on the Prevention of Escherichia coli Biofilm Formation and Susceptibility." Antibiotics 9, no. 5 (April 29, 2020): 216. http://dx.doi.org/10.3390/antibiotics9050216.

Full text
Abstract:
Urinary tract infections are one of the most common hospital-acquired infections, and they are often associated with biofilm formation in indwelling medical devices such as catheters and stents. This study aims to investigate the antibiofilm performance of a polymer brush—poly[oligo(ethylene glycol) methyl ether methacrylate], poly(MeOEGMA)—and evaluate its effect on the antimicrobial susceptibility of Escherichia coli biofilms formed on that surface. Biofilms were formed in a parallel plate flow chamber (PPFC) for 24 h under the hydrodynamic conditions prevailing in urinary catheters and stents and challenged with ampicillin. Results obtained with the brush were compared to those obtained with two control surfaces, polydimethylsiloxane (PDMS) and glass. The polymer brush reduced by 57% the surface area covered by E. coli after 24 h, as well as the number of total adhered cells. The antibiotic treatment potentiated cell death and removal, and the total cell number was reduced by 88%. Biofilms adapted their architecture, and cell morphology changed to a more elongated form during that period. This work suggests that the poly(MeOEGMA) brush has potential to prevent bacterial adhesion in urinary tract devices like ureteral stents and catheters, as well as in eradicating biofilms developed in these biomedical devices.
APA, Harvard, Vancouver, ISO, and other styles
7

Liu, Caihong, Jongho Lee, Jun Ma, and Menachem Elimelech. "Antifouling Thin-Film Composite Membranes by Controlled Architecture of Zwitterionic Polymer Brush Layer." Environmental Science & Technology 51, no. 4 (February 2, 2017): 2161–69. http://dx.doi.org/10.1021/acs.est.6b05992.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Yaremchuk, D., T. Patsahan, and J. Ilnytskyi. "Photo-switchable liquid crystalline brush as an aligning surface for liquid crystals: modelling via mesoscopic computer simulations." Condensed Matter Physics 25, no. 3 (2022): 33601. http://dx.doi.org/10.5488/cmp.25.33601.

Full text
Abstract:
We consider the mesoscopic model for the liquid crystalline brush that might serve as a photoswitchable aligning surface for preorientation of low molecular weight liquid crystals in a bulk. The brush is built by grafting the polymer chains of a side-chain molecular architecture, with the side chains terminated by a chromophore unit mimicking the azobenzene unit, to a substrate. When irradiated with ultraviolet light, the chromophores photoisomerize into a non-mesogenic cis state and the whole system turns into an ordinary polymer brush with no orientational order and two states: the collapsed and straightened one, depending on the grafting density. When irradiated with visible light, the chromophores photoisomerize into a mesogenic trans state, resulting in formation of a transient network between chains because of a strong attraction between chromophores. Spontaneous self-assembly of the brush in these conditions results in an orientationally isotropic polydomain structure. The desired uniaxial planar ordering of chromophores within a brush can be achieved at certain temperature and grafting density intervals, as the result of a two-stage preparation protocol. An external stimulus orients chromophores uniaxially at the first stage. The system is equilibrated at the second stage at a given temperature and with the external stimulus switched off. The preoriented chromophores either keep or loose their orientations depending on the strength of the memory effect inherent to a transient network of chains that are formed during the first stage, similarly to the case of the liquid crystalline elastomers, where such effects are caused by the covalent crosslinks.
APA, Harvard, Vancouver, ISO, and other styles
9

Laktionov, Mikhail Y., Ekaterina B. Zhulina, Ralf P. Richter, and Oleg V. Borisov. "Polymer Brush in a Nanopore: Effects of Solvent Strength and Macromolecular Architecture Studied by Self-Consistent Field and Scaling Theory." Polymers 13, no. 22 (November 14, 2021): 3929. http://dx.doi.org/10.3390/polym13223929.

Full text
Abstract:
To study conformational transition occuring upon inferior solvent strength in a brush formed by linear or dendritically branched macromolecules tethered to the inner surface of cylindrical or planar (slit-like) pore, a self-consistent field analytical approach is employed. Variations in the internal brush structure as a function of variable solvent strength and pore radius, and the onset of formation of a hollow channel in the pore center are analysed. The predictions of analytical theory are supported and complemented by numerical modelling by a self-consistent field Scheutjens–Fleer method. Scaling arguments are used to study microphase segregation under poor solvent conditions leading to formation of a laterally and longitudinally patterned structure in planar and cylindrical pores, respectively, and the effects of confinement on "octopus-like" clusters in the pores of different geometries.
APA, Harvard, Vancouver, ISO, and other styles
10

Wang, Gang, Wei Huang, Nicholas D. Eastham, Simone Fabiano, Eric F. Manley, Li Zeng, Binghao Wang, et al. "Aggregation control in natural brush-printed conjugated polymer films and implications for enhancing charge transport." Proceedings of the National Academy of Sciences 114, no. 47 (November 6, 2017): E10066—E10073. http://dx.doi.org/10.1073/pnas.1713634114.

Full text
Abstract:
Shear-printing is a promising processing technique in organic electronics for microstructure/charge transport modification and large-area film fabrication. Nevertheless, the mechanism by which shear-printing can enhance charge transport is not well-understood. In this study, a printing method using natural brushes is adopted as an informative tool to realize direct aggregation control of conjugated polymers and to investigate the interplay between printing parameters, macromolecule backbone alignment and aggregation, and charge transport anisotropy in a conjugated polymer series differing in architecture and electronic structure. This series includes (i) semicrystalline hole-transporting P3HT, (ii) semicrystalline electron-transporting N2200, (iii) low-crystallinity hole-transporting PBDTT-FTTE, and (iv) low-crystallinity conducting PEDOT:PSS. The (semi-)conducting films are characterized by a battery of morphology and microstructure analysis techniques and by charge transport measurements. We report that remarkably enhanced mobilities/conductivities, as high as 5.7×/3.9×, are achieved by controlled growth of nanofibril aggregates and by backbone alignment, with the adjusted R2 (R2adj) correlation between aggregation and charge transport as high as 95%. However, while shear-induced aggregation is important for enhancing charge transport, backbone alignment alone does not guarantee charge transport anisotropy. The correlations between efficient charge transport and aggregation are clearly shown, while mobility and degree of orientation are not always well-correlated. These observations provide insights into macroscopic charge transport mechanisms in conjugated polymers and suggest guidelines for optimization.
APA, Harvard, Vancouver, ISO, and other styles
11

Uekusa, Takayuki, Shusaku Nagano, and Takahiro Seki. "Highly Ordered In-Plane Photoalignment Attained by the Brush Architecture of Liquid Crystalline Azobenzene Polymer." Macromolecules 42, no. 1 (January 13, 2009): 312–18. http://dx.doi.org/10.1021/ma802010x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Lee, Dong Hoon, Yoko Tokuno, Satoshi Uchida, Masaaki Ozawa, and Koji Ishizu. "Architecture of polymer particles composed of brush structure at surfaces and construction of colloidal crystals." Journal of Colloid and Interface Science 340, no. 1 (December 2009): 27–34. http://dx.doi.org/10.1016/j.jcis.2009.08.035.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Sun, Zhexun, Elizabeth Feeney, Ya Guan, Sierra G. Cook, Delphine Gourdon, Lawrence J. Bonassar, and David Putnam. "Boundary mode lubrication of articular cartilage with a biomimetic diblock copolymer." Proceedings of the National Academy of Sciences 116, no. 25 (June 4, 2019): 12437–41. http://dx.doi.org/10.1073/pnas.1900716116.

Full text
Abstract:
We report the design of a diblock copolymer with architecture and function inspired by the lubricating glycoprotein lubricin. This diblock copolymer, synthesized by sequential reversible addition–fragmentation chain-transfer polymerization, consists of a cationic cartilage-binding domain and a brush-lubricating domain. It reduces the coefficient of friction of articular cartilage under boundary mode conditions (0.088 ± 0.039) to a level equivalent to that provided by lubricin (0.093 ± 0.011). Additionally, both the EC50 (0.404 mg/mL) and cartilage-binding time constant (7.19 min) of the polymer are comparable to purified human and recombinant lubricin. Like lubricin, the tribological properties of this polymer are dependent on molecular architecture. When the same monomer composition was evaluated either as an AB diblock copolymer or as a random copolymer, the diblock effectively lubricated cartilage under boundary mode conditions whereas the random copolymer did not. Additionally, the individual polymer blocks did not lubricate independently, and lubrication could be competitively inhibited with an excess of binding domain. This diblock copolymer is an example of a synthetic polymer with lubrication properties equal to lubricin under boundary mode conditions, suggesting its potential utility as a therapy for joint pathologies like osteoarthritis.
APA, Harvard, Vancouver, ISO, and other styles
14

Sidoli, Ugo, Hisaschi T. Tee, Ivan Raguzin, Jakob Mühldorfer, Frederik R. Wurm, and Alla Synytska. "Thermo-Responsive Polymer Brushes with Side Graft Chains: Relationship Between Molecular Architecture and Underwater Adherence." International Journal of Molecular Sciences 20, no. 24 (December 13, 2019): 6295. http://dx.doi.org/10.3390/ijms20246295.

Full text
Abstract:
During the last few decades, wet adhesives have been developed for applications in various fields. Nonetheless, key questions such as the most suitable polymer architecture as well as the most suitable chemical composition remain open. In this article, we investigate the underwater adhesion properties of novel responsive polymer brushes with side graft chain architecture prepared using “grafting through” approach on flat surfaces. The incorporation in the backbone of thermo-responsive poly(N-isopropylacrylamide) (PNIPAm) allowed us to obtain LCST behavior in the final layers. PNIPAm is co-polymerized with poly(methyl ethylene phosphate) (PMEP), a poloyphosphoester. The final materials are characterized studying the surface-grafted polymer as well as the polymer from the bulk solution, and pure PNIPAm brush is used as reference. PNIPAm-g-PMEP copolymers retain the responsive behavior of PNIPAm: when T > LCST, a clear switching of properties is observed. More specifically, all layers above the critical temperature show collapse of the chains, increased hydrophobicity and variation of the surface charge even if no ionizable groups are present. Secondly, effect of adhesion parameters such as debonding rate and contact time is studied. Thirdly, the reversibility of the adhesive properties is confirmed by performing adhesion cycles. Finally, the adhesive properties of the layers are studied below and above the LCST against hydrophilic and hydrophobic substrates.
APA, Harvard, Vancouver, ISO, and other styles
15

Dang, Alei, Chin Ming Hui, Rachel Ferebee, Joshua Kubiak, Tiehu Li, Krzysztof Matyjaszewski, and Michael R. Bockstaller. "Thermal Properties of Particle Brush Materials: Effect of Polymer Graft Architecture on the Glass Transition Temperature in Polymer-Grafted Colloidal Systems." Macromolecular Symposia 331-332, no. 1 (October 2013): 9–16. http://dx.doi.org/10.1002/masy.201300062.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Li, Danyang, Momina Ahmed, Anisah Khan, Lizhou Xu, Adam A. Walters, Belén Ballesteros, and Khuloud T. Al-Jamal. "Tailoring the Architecture of Cationic Polymer Brush-Modified Carbon Nanotubes for Efficient siRNA Delivery in Cancer Immunotherapy." ACS Applied Materials & Interfaces 13, no. 26 (June 25, 2021): 30284–94. http://dx.doi.org/10.1021/acsami.1c02627.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Ishizu, Koji, Isamu Amir, Nobuyuki Okamoto, Satoshi Uchida, Masaaki Ozawa, and Hui Chen. "Synthesis of tailored core–brush polymer particles via a living radical polymerization and architecture of colloidal crystals." Journal of Colloid and Interface Science 353, no. 1 (January 2011): 69–75. http://dx.doi.org/10.1016/j.jcis.2010.08.067.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Olejnik, Piotr, Marianna Gniadek, Luis Echegoyen, and Marta Plonska-Brzezinska. "Nanoforest: Polyaniline Nanotubes Modified with Carbon Nano-Onions as a Nanocomposite Material for Easy-to-Miniaturize High-Performance Solid-State Supercapacitors." Polymers 10, no. 12 (December 19, 2018): 1408. http://dx.doi.org/10.3390/polym10121408.

Full text
Abstract:
This article describes a facile low-cost synthesis of polyaniline nanotube (PANINT)–carbon nano-onion (CNO) composites for solid-state supercapacitors. Scanning electron microscopic (SEM) analyses indicate a uniform and ordered composition for the conducting polymer nanotubes immobilized on a thin gold film. The obtained nanocomposites exhibit a brush-like architecture with a specific capacitance of 946 F g−1 at a scan rate of 1 mV s−1. In addition, the nanocomposites offer high conductivity and a porous and well-developed surface area. The PANINT–CNO nanocomposites were tested as electrodes with high potential and long-term stability for use in easy-to-miniaturize high-performance supercapacitor devices.
APA, Harvard, Vancouver, ISO, and other styles
19

Wang, Dali, Jiaqi Lin, Fei Jia, Xuyu Tan, Yuyan Wang, Xiaoya Sun, Xueyan Cao, et al. "Bottlebrush-architectured poly(ethylene glycol) as an efficient vector for RNA interference in vivo." Science Advances 5, no. 2 (February 2019): eaav9322. http://dx.doi.org/10.1126/sciadv.aav9322.

Full text
Abstract:
Nonhepatic delivery of small interfering RNAs (siRNAs) remains a challenge for development of RNA interference–based therapeutics. We report a noncationic vector wherein linear poly(ethylene glycol) (PEG), a polymer generally considered as inert and safe biologically but ineffective as a vector, is transformed into a bottlebrush architecture. This topology provides covalently embedded siRNA with augmented nuclease stability and cellular uptake. Consisting almost entirely of PEG and siRNA, the conjugates exhibit a ~25-fold increase in blood elimination half-life and a ~19-fold increase in the area under the curve compared with unmodified siRNA. The improved pharmacokinetics results in greater tumor uptake and diminished liver capture. Despite the structural simplicity these conjugates efficiently knock down target genes in vivo without apparent toxic and immunogenic reactions. Given the benign biological nature of PEG and its widespread precedence in biopharmaceuticals, we anticipate the brush polymer–based technology to have a significant impact on siRNA therapeutics.
APA, Harvard, Vancouver, ISO, and other styles
20

Schmitt, Michael, Jianan Zhang, Jaejun Lee, Bongjoon Lee, Xin Ning, Ren Zhang, Alamgir Karim, Robert F. Davis, Krzysztof Matyjaszewski, and Michael R. Bockstaller. "Polymer ligand–induced autonomous sorting and reversible phase separation in binary particle blends." Science Advances 2, no. 12 (December 2016): e1601484. http://dx.doi.org/10.1126/sciadv.1601484.

Full text
Abstract:
The tethering of ligands to nanoparticles has emerged as an important strategy to control interactions and organization in particle assembly structures. We demonstrate that ligand interactions in mixtures of polymer-tethered nanoparticles (which are modified with distinct types of polymer chains) can impart upper or lower critical solution temperature (UCST/LCST)–type phase behavior on binary particle mixtures in analogy to the phase behavior of the corresponding linear polymer blends. Therefore, cooling (or heating) of polymer-tethered particle blends with appropriate architecture to temperatures below (or above) the UCST (or LCST) results in the organization of the individual particle constituents into monotype microdomain structures. The shape (bicontinuous or island-type) and lengthscale of particle microdomains can be tuned by variation of the composition and thermal process conditions. Thermal cycling of LCST particle brush blends through the critical temperature enables the reversible growth and dissolution of monoparticle domain structures. The ability to autonomously and reversibly organize multicomponent particle mixtures into monotype microdomain structures could enable transformative advances in the high-throughput fabrication of solid films with tailored and mutable structures and properties that play an important role in a range of nanoparticle-based material technologies.
APA, Harvard, Vancouver, ISO, and other styles
21

Totani, Masayasu, Tsuyoshi Ando, Kayo Terada, Takaya Terashima, Ill Yong Kim, Chikara Ohtsuki, Chuanwu Xi, Kenichi Kuroda, and Masao Tanihara. "Utilization of star-shaped polymer architecture in the creation of high-density polymer brush coatings for the prevention of platelet and bacteria adhesion." Biomaterials Science 2, no. 9 (May 22, 2014): 1172. http://dx.doi.org/10.1039/c4bm00034j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Shibuya, Yoshiki, Ryoichi Tatara, Yivan Jiang, Yang Shao‐Horn, and Jeremiah A. Johnson. "Brush‐First ROMP of poly(ethylene oxide) macromonomers of varied length: impact of polymer architecture on thermal behavior and Li + conductivity." Journal of Polymer Science Part A: Polymer Chemistry 57, no. 3 (October 9, 2018): 448–55. http://dx.doi.org/10.1002/pola.29242.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Al Nakeeb, Nischang, and Schmidt. "Tannic Acid-Mediated Aggregate Stabilization of Poly(N-vinylpyrrolidone)-b-poly(oligo (ethylene glycol) methyl ether methacrylate) Double Hydrophilic Block Copolymers." Nanomaterials 9, no. 5 (April 26, 2019): 662. http://dx.doi.org/10.3390/nano9050662.

Full text
Abstract:
The self-assembly of block copolymers in aqueous solution is an important field in modern polymer science that has been extended to double hydrophilic block copolymers (DHBC) in recent years. In here, a significant improvement of the self-assembly process of DHBC in aqueous solution by utilizing a linear-brush macromolecular architecture is presented. The improved self-assembly behavior of poly(N-vinylpyrrolidone)-b-poly(oligo(ethylene glycol) methyl ether methacrylate) (PVP-b-P(OEGMA)) and its concentration dependency is investigated via dynamic light scattering (DLS) (apparent hydrodynamic radii ≈ 100–120 nm). Moreover, the DHBC assemblies can be non-covalently crosslinked with tannic acid via hydrogen bonding, which leads to the formation of small aggregates as well (apparent hydrodynamic radius ≈ 15 nm). Non-covalent crosslinking improves the self-assembly and stabilizes the aggregates upon dilution, reducing the concentration dependency of aggregate self-assembly. Additionally, the non-covalent aggregates can be disassembled in basic media. The presence of aggregates was studied via cryogenic scanning electron microscopy (cryo-SEM) and DLS before and after non-covalent crosslinking. Furthermore, analytical ultracentrifugation of the formed aggregate structures was performed, clearly showing the existence of polymer assemblies, particularly after non-covalent crosslinking. In summary, we report on the completely hydrophilic self-assembled structures in solution formed from fully biocompatible building entities in water.
APA, Harvard, Vancouver, ISO, and other styles
24

Saha, Debasish, Karthik R. Peddireddy, Jürgen Allgaier, Wei Zhang, Simona Maccarrone, Henrich Frielinghaus, and Dieter Richter. "Amphiphilic Comb Polymers as New Additives in Bicontinuous Microemulsions." Nanomaterials 10, no. 12 (December 2, 2020): 2410. http://dx.doi.org/10.3390/nano10122410.

Full text
Abstract:
It has been shown that the thermodynamics of bicontinuous microemulsions can be tailored via the addition of various different amphiphilic polymers. In this manuscript, we now focus on comb-type polymers consisting of hydrophobic backbones and hydrophilic side chains. The distinct philicity of the backbone and side chains leads to a well-defined segregation into the oil and water domains respectively, as confirmed by contrast variation small-angle neutron scattering experiments. This polymer–microemulsion structure leads to well-described conformational entropies of the polymer fragments (backbone and side chains) that exert pressure on the membrane, which influences the thermodynamics of the overall microemulsion. In the context of the different polymer architectures that have been studied by our group with regards to their phase diagrams and small-angle neutron scattering, the microemulsion thermodynamics of comb polymers can be described in terms of a superposition of the backbone and side chain fragments. The denser or longer the side chain, the stronger the grafting and the more visible the brush effect of the side chains becomes. Possible applications of the comb polymers as switchable additives are discussed. Finally, a balanced philicity of polymers also motivates transmembrane migration in biological systems of the polymers themselves or of polymer–DNA complexes.
APA, Harvard, Vancouver, ISO, and other styles
25

Reese, Cassandra M., Brittany J. Thompson, Phillip K. Logan, Christopher M. Stafford, Michael Blanton, and Derek L. Patton. "Sequential and one-pot post-polymerization modification reactions of thiolactone-containing polymer brushes." Polymer Chemistry 10, no. 36 (2019): 4935–43. http://dx.doi.org/10.1039/c9py01123d.

Full text
Abstract:
Polymer brushes carrying pendent thiolactone functional groups were explored for the design of multifunctional homopolymer brush architectures using sequential and one-pot postpolymerization strategies.
APA, Harvard, Vancouver, ISO, and other styles
26

Sun, Dachuan, and Yang Song. "Influences of the Periodicity in Molecular Architecture on the Phase Diagrams and Microphase Transitions of the Janus Double-Brush Copolymer with a Loose Graft." Polymers 14, no. 14 (July 13, 2022): 2847. http://dx.doi.org/10.3390/polym14142847.

Full text
Abstract:
The backbone of the Janus double-brush copolymer may break during long-term service, but whether this breakage affects the self-assembled phase state and microphase transitions of the material is still unknown. For the Janus double-brush copolymers with a periodicity in molecular architecture ranging from 1 to 10, the influences of the architectural periodicity on their phase diagrams and order–disorder transitions (ODT) were investigated by the self-consistent mean field theory (SCFT). In total, nine microphases with long-range order were found. By comparing the phase diagrams between copolymers of different periodicity, a decrease in periodicity or breakage along the copolymer backbone had nearly no influence on the phase diagrams unless the periodicity was too short to be smaller than 3. For copolymers with neutral backbones, a decrease in periodicity or breakage along the copolymer backbone reduced the critical segregation strengths of the whole copolymer at ODT. The equations for the critical segregation strengths at ODT, the architectural periodicity, and the volume fraction of the backbone were established for the Janus double-brush copolymers. The theoretical calculations were consistent with the previous theoretical, experimental, and simulation results.
APA, Harvard, Vancouver, ISO, and other styles
27

Besford, Quinn A., Simon Schubotz, Soosang Chae, Ayşe B. Özdabak Sert, Alessia C. G. Weiss, Günter K. Auernhammer, Petra Uhlmann, José Paulo S. Farinha, and Andreas Fery. "Molecular Transport within Polymer Brushes: A FRET View at Aqueous Interfaces." Molecules 27, no. 9 (May 9, 2022): 3043. http://dx.doi.org/10.3390/molecules27093043.

Full text
Abstract:
Molecular permeability through polymer brush chains is implicated in surface lubrication, wettability, and solute capture and release. Probing molecular transport through polymer brushes can reveal information on the polymer nanostructure, with a permeability that is dependent on chain conformation and grafting density. Herein, we introduce a brush system to study the molecular transport of fluorophores from an aqueous droplet into the external “dry” polymer brush with the vapour phase above. The brushes consist of a random copolymer of N-isopropylacrylamide and a Förster resonance energy transfer (FRET) donor-labelled monomer, forming ultrathin brush architectures of about 35 nm in solvated height. Aqueous droplets containing a separate FRET acceptor are placed onto the surfaces, with FRET monitored spatially around the 3-phase contact line. FRET is used to monitor the transport from the droplet to the outside brush, and the changing internal distributions with time as the droplets prepare to recede. This reveals information on the dynamics and distances involved in the molecular transport of the FRET acceptor towards and away from the droplet contact line, which are strongly dependent on the relative humidity of the system. We anticipate our system to be extremely useful for studying lubrication dynamics and surface droplet wettability processes.
APA, Harvard, Vancouver, ISO, and other styles
28

Gowneni, Soujanya, Kota Ramanjaneyulu, and Pratyay Basak. "Polymer-Nanocomposite Brush-like Architectures as an All-Solid Electrolyte Matrix." ACS Nano 8, no. 11 (November 13, 2014): 11409–24. http://dx.doi.org/10.1021/nn504472v.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Altay, Esra, and Javid Rzayev. "Synthesis of star-brush polymer architectures from end-reactive molecular bottlebrushes." Polymer 98 (August 2016): 487–94. http://dx.doi.org/10.1016/j.polymer.2016.02.022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Gadzinowski, Mariusz, Maciej Kasprów, Teresa Basinska, Stanislaw Slomkowski, Łukasz Otulakowski, Barbara Trzebicka, and Tomasz Makowski. "Synthesis, Hydrophilicity and Micellization of Coil-Brush Polystyrene-b-(polyglycidol-g-polyglycidol) Copolymer—Comparison with Linear Polystyrene-b-polyglycidol." Polymers 14, no. 2 (January 8, 2022): 253. http://dx.doi.org/10.3390/polym14020253.

Full text
Abstract:
In this paper, an original method of synthesis of Coil-Brush amphiphilic polystyrene-b-(polyglycidol-g-polyglycidol) (PS-b-(PGL-g-PGL)) block copolymers was developed. The hypothesis that their hydrophilicity and micellization can be controlled by polyglycidol blocks architecture was verified. The research enabled comparison of behavior in water of PS-b-PGL copolymers and block–brush copolymers PS-b-(PGL-g-PGL) with similar composition. The Coil-Brush copolymers were composed of PS-b-PGL linear core with average DPn of polystyrene 29 and 13 of polyglycidol blocks. The DPn of polyglycidol side blocks of coil–b–brush copolymers were 2, 7, and 11, respectively. The copolymers were characterized by 1H and 13C NMR, GPC, and FTIR methods. The hydrophilicity of films from the linear and Coil-Brush copolymers was determined by water contact angle measurements in static conditions. The behavior of Coil-Brush copolymers in water and their critical micellization concentration (CMC) were determined by UV-VIS using 1,6-diphenylhexa-1,3,5-trien (DPH) as marker and by DLS. The CMC values for brush copolymers were much higher than for linear species with similar PGL content. The results of the copolymer film wettability and the copolymer self-assembly studies were related to fraction of hydrophilic polyglycidol. The CMC for both types of polymers increased exponentially with increasing content of polyglycidol.
APA, Harvard, Vancouver, ISO, and other styles
31

Hsu, Hsiao-Ping, and Wolfgang Paul. "A fast Monte Carlo algorithm for studying bottle-brush polymers." Computer Physics Communications 182, no. 10 (October 2011): 2115–21. http://dx.doi.org/10.1016/j.cpc.2011.05.005.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Neratova, Irina V., Torsten Kreer, and Jens-Uwe Sommer. "Translocation of Molecules with Different Architectures through a Brush-Covered Microchannel." Macromolecules 48, no. 11 (May 7, 2015): 3756–66. http://dx.doi.org/10.1021/acs.macromol.5b00042.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Wu, Tong, Xuefei Leng, Yanshai Wang, Zhiyong Wei, and Yang Li. "Linear- and star-brush poly(ethylene glycol)s: Synthesis and architecture-dependent crystallization behavior." Polymer 202 (August 2020): 122661. http://dx.doi.org/10.1016/j.polymer.2020.122661.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Wang, Shi, Lei Zhang, Qinghui Zeng, Xu Liu, Wen-Yong Lai, and Liaoyun Zhang. "Cellulose Microcrystals with Brush-Like Architectures as Flexible All-Solid-State Polymer Electrolyte for Lithium-Ion Battery." ACS Sustainable Chemistry & Engineering 8, no. 8 (February 10, 2020): 3200–3207. http://dx.doi.org/10.1021/acssuschemeng.9b06658.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Yan, Jiajun, Michael R. Bockstaller, and Krzysztof Matyjaszewski. "Brush-modified materials: Control of molecular architecture, assembly behavior, properties and applications." Progress in Polymer Science 100 (January 2020): 101180. http://dx.doi.org/10.1016/j.progpolymsci.2019.101180.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Maw, Mitchell, Benjamin J. Morgan, Erfan Dashtimoghadam, Yuan Tian, Egor A. Bersenev, Alina V. Maryasevskaya, Dimitri A. Ivanov, Krzysztof Matyjaszewski, Andrey V. Dobrynin, and Sergei S. Sheiko. "Brush Architecture and Network Elasticity: Path to the Design of Mechanically Diverse Elastomers." Macromolecules 55, no. 7 (March 29, 2022): 2940–51. http://dx.doi.org/10.1021/acs.macromol.2c00006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Wang, Zhanhua, and Han Zuilhof. "Antifouling Properties of Fluoropolymer Brushes toward Organic Polymers: The Influence of Composition, Thickness, Brush Architecture, and Annealing." Langmuir 32, no. 26 (June 22, 2016): 6571–81. http://dx.doi.org/10.1021/acs.langmuir.6b00695.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Lyu, Yu-Feng, Zhi-Jie Zhang, Chang Liu, Zhi Geng, Long-Cheng Gao, and Quan Chen. "Random binary brush architecture enhances both ionic conductivity and mechanical strength at room temperature." Chinese Journal of Polymer Science 36, no. 1 (November 3, 2017): 78–84. http://dx.doi.org/10.1007/s10118-018-2016-z.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Gon, Saugata, and Maria M. Santore. "Sensitivity of Protein Adsorption to Architectural Variations in a Protein-Resistant Polymer Brush Containing Engineered Nanoscale Adhesive Sites." Langmuir 27, no. 24 (December 20, 2011): 15083–91. http://dx.doi.org/10.1021/la203293k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Hong, Jisu, Yonghwan Kwon, Min Sang Kwon, and Chaenyung Cha. "Aziridine-Capped Poly(ethylene glycol) Brush Copolymers with Tunable Architecture as Versatile Cross-Linkers for Adhesives." ACS Applied Polymer Materials 4, no. 3 (February 25, 2022): 2105–15. http://dx.doi.org/10.1021/acsapm.1c01894.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Ishizu, Koji, Naomasa Hatoyama, and Satoshi Uchida. "Architecture of rod–brush block copolymers synthesized by a combination of coordination polymerization and atom transfer radical polymerization." Journal of Applied Polymer Science 108, no. 5 (2008): 3346–52. http://dx.doi.org/10.1002/app.27936.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Fei, Hua-Feng, Benjamin M. Yavitt, Xiyu Hu, Gayathri Kopanati, Alexander Ribbe, and James J. Watkins. "Influence of Molecular Architecture and Chain Flexibility on the Phase Map of Polystyrene-block-poly(dimethylsiloxane) Brush Block Copolymers." Macromolecules 52, no. 17 (August 21, 2019): 6449–57. http://dx.doi.org/10.1021/acs.macromol.9b00843.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Theodosopoulos, George V., Spyridoula-Lida Bitsi, and Marinos Pitsikalis. "Complex Brush-Like Macromolecular Architectures via Anionic and Ring Opening Metathesis Polymerization: Synthesis, Characterization, and Thermal Properties." Macromolecular Chemistry and Physics 219, no. 1 (September 27, 2017): 1700253. http://dx.doi.org/10.1002/macp.201700253.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Zhu, Linlin, Huishan Huang, Ying Wang, Zhen Zhang, and Nikos Hadjichristidis. "Organocatalytic Synthesis of Polysulfonamides with Well-Defined Linear and Brush Architectures from a Designed/Synthesized Bis(N-sulfonyl aziridine)." Macromolecules 54, no. 17 (August 11, 2021): 8164–72. http://dx.doi.org/10.1021/acs.macromol.1c01193.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Orobets, Julian, and Oleh Rybchynskyy. "THE RESTORATION PROGRAM OF THE WHITESTONE SCULPTURE OF THE VIRGIN OF THE IMMACULATE CONCEPTION FROM THE VILLAGE OF DOBRYANY HORODOK DISTRICT." Current Issues in Research, Conservation and Restoration of Historic Fortifications 16, no. 2022 (2022): 83–90. http://dx.doi.org/10.23939/fortifications2022.16.083.

Full text
Abstract:
The article demonstrates the peculiarities of the artistic sculpture of the Virgin Mary in the sculpture of the late XIX-first half of the XX century. The main positions of the restoration program and the course of its implementation are revealed. The plastic nature of the folds and compositional principles suggest that the sculptor worked in the mid-XIX - early XX centuries. The textural and petrographic properties of the stone are the basis for the assumption that the author worked in the village of Demnya because the complex of the church of St. Nicholas and the one analyzed in the article have many similarities. Comparing some of the sculptures of famous authors of the late Baroque and Classicism, we can identify fundamental similarities: portrait features, execution of tears, and the nature of the image of the lips, feet and hands. In the sculpture of the Mother of God from the village of Dobryany, Horodok district, the influence of the works of Peter Viitovych, Leonardo Marconi, Vuytsyk, Plishevsky and Soltys is noticeable. The works of these authors are characterized by realism, careful expression of the anatomy of the human body, true depiction of movement and clear composition. Before the restoration work began, the white stone sculpture was in poor condition. Below the figure, on a memorial plaque, it is written that masters from the village of Demnya restored it in 2000. 5 layers of whitewash were found on the sculpture. The sculpture in the village was whitewashed with lime in honour of holidays and solemn events. Initially, the figure was not covered with paint. The figure of the Virgin was divided into two parts. The lower bullet with the snake and the feet was broken off from the rest of the body. These parts were previously fastened to an anchor made of black iron and cement. Anchor slashed hard and spread the stone. As a result, the folds of the dress and previous masterpieces were peeled off at the bottom. After a thorough examination, non-professional fastening of the folded arms was revealed. There were inaccurate traces of cement additions and cracks in the places of glueing. A profiled base was also glued to the ball on a dense layer of cement. In general, due to the colonies of bio-growers and large areas of whitewash, the work lost its aesthetic appeal and symbolic content. The first stage of restoration consisted of the layer-by-layer sounding of limestone. After opening the limestone, the object was dry cleaned. Dry cleaning was done with scalpels, nylon brushes, and small brass and steel brushes, not durable places were tapped with a chisel in order not to damage the original surface of the stone. When the lime filler was poorly removed, the method of wet cleaning was used, a detergent solution dissolved the lime filler. The detergent solution was applied with a flute brush and the fluffy surface was cleaned with a nylon brush. The detergent solution is made for better penetration of surfactants into the pores - soap softens the surface of the water, which helps to dissolve the dirt in the pores of the stone. Subsequently, the stage of structural strengthening of the stone was performed. To achieve this goal, the organosilicon hardener Remmers KSE 300 was used. The solution was applied with a brush. To crystallize the organosilicon hardener a technological break was taken for two weeks, moving on to the stage of glueing parts. Bonding of elements took place with the help of epoxy glue from TENAX. Before applying the glue, the elements were coated with a 3% alcohol solution of polymer "Paraloid B72". The upper and lower parts of the sculpture were attached to epoxy resin and two stainless steel rods. The lost fragments of the folds of the Virgin's clothing were made of fibreglass reinforcement attached to the holes of the stone with epoxy glue. After completing the stage of glueing the sculpture, the addition of lost elements was performed. A mineral carbonate solution was used. After a two-week technological break, the stage of toning the supplemented parts began in order to achieve organic homogeneity of the figure. The next step is to cover the stone with a long-acting biocidal solution. At the end of the restoration, the work should be covered with lime water to even out the overall tone of the sculpture. The sculptural composition of the Immaculate Conception of the Virgin Mary from the village of Dobryany, Horodok district, is of great artistic and historical value. During the restoration of the sculpture the compatible materials were used as well as generally accepted and original restoration tools, which will allow the exhibition of the work of art in an authentic place near the chapel., After returning the work to its holders, it is recommended to exhibit it under an architectural cover to prevent the aggressive effects of precipitation and the negative changes in temperature and humidity in autumn-winter.
APA, Harvard, Vancouver, ISO, and other styles
46

Jing, Benxin, Xiaofeng Wang, Yi Shi, Yingxi Zhu, Haifeng Gao, and Susan K. Fullerton-Shirey. "Combining Hyperbranched and Linear Structures in Solid Polymer Electrolytes to Enhance Mechanical Properties and Room-Temperature Ion Transport." Frontiers in Chemistry 9 (June 25, 2021). http://dx.doi.org/10.3389/fchem.2021.563864.

Full text
Abstract:
Polyethylene oxide (PEO)-based polymers are commonly studied for use as a solid polymer electrolyte for rechargeable Li-ion batteries; however, simultaneously achieving sufficient mechanical integrity and ionic conductivity has been a challenge. To address this problem, a customized polymer architecture is demonstrated wherein PEO bottle-brush arms are hyperbranched into a star architecture and then functionalized with end-grafted, linear PEO chains. The hierarchical architecture is designed to minimize crystallinity and therefore enhance ion transport via hyperbranching, while simultaneously addressing the need for mechanical integrity via the grafting of long, PEO chains (Mn = 10,000). The polymers are doped with lithium bis(trifluoromethane) sulfonimide (LiTFSI), creating hierarchically hyperbranched (HB) solid polymer electrolytes. Compared to electrolytes prepared with linear PEO of equivalent molecular weight, the HB PEO electrolytes increase the room temperature ionic conductivity from ∼2.5 × 10–6 to 2.5 × 10−5 S/cm. The conductivity increases by an additional 50% by increasing the block length of the linear PEO in the bottle brush arms from Mn = 1,000 to 2,000. The mechanical properties are improved by end-grafting linear PEO (Mn = 10,000) onto the terminal groups of the HB PEO bottle-brush. Specifically, the Young’s modulus increases by two orders of magnitude to a level comparable to commercial PEO films, while only reducing the conductivity by 50% below the HB electrolyte without grafted PEO. This study addresses the trade-off between ion conductivity and mechanical properties, and shows that while significant improvements can be made to the mechanical properties with hierarchical grafting of long, linear chains, only modest gains are made in the room temperature conductivity.
APA, Harvard, Vancouver, ISO, and other styles
47

Onoda, Michika, Fei Jia, Yukikazu Takeoka, and Robert Macfarlane. "Controlling the dynamics of elastomer networks with multivalent brush architectures." Soft Matter, 2022. http://dx.doi.org/10.1039/d2sm00328g.

Full text
Abstract:
Herein, we report a design strategy for developing mechanically enhanced and dynamic polymer networks by incorporating a polymer with multivalent brush architecture. Different ratios of two types of imidazole functionalized...
APA, Harvard, Vancouver, ISO, and other styles
48

Gersappe, Dilip, Michael Fasolka, Rafel Israels, and Anna C. Balazs. "Tailoring the Structure of Polymer Brushes Through Copolymer Architecture." MRS Proceedings 385 (1995). http://dx.doi.org/10.1557/proc-385-201.

Full text
Abstract:
ABSTRACTPolymers tethered by one end onto a solid surface are referred to as polymer “brushes”. We consider brushes composed of copolymers that contain both A and B monomers. The A monomers are compatible with the surrounding solvent, while the B sites are solventincompatible. The solvent incompatibility causes the B sites to associate into domains or clusters within the layer. We use Monte Carlo computer simulations and self-consistent field calculations to determine the effect of copolymer architecture on the structure of the polymer brush. In particular, we alter the copolymer sequence distribution (the arrangement of the A and B monomers along the length of the chain) and determine how both the vertical and lateral morphology of the brush are effected by these variations. The results provide guidelines for controlling the size and shape of the B domains, and consequently, the morphology of the tethered layer.
APA, Harvard, Vancouver, ISO, and other styles
49

Sawhney, U., C. J. Durning, B. O'Shaughnessy, G. S. Smith, and J. Majewski. "Irreversible Adsorption of Polymer Melts." MRS Proceedings 464 (1996). http://dx.doi.org/10.1557/proc-464-219.

Full text
Abstract:
ABSTRACTWe studied the equilibrium architecture of polymer layers strongly adsorbed from the melt. Immobilized layers of poly-(methyl methacrylate) (PMMA) were produced by the following method: 1) The polymer was spin-coated onto silanol bearing surfaces of single crystal and fused quartz, and annealed at melt conditions, 2) The annealed layer was quenched to room temperature (below the glass transition temperature) in order to “freeze in” the melt structure near the substrate, 3) Unbound material was leached away in good solvent (benzene) to leave a residual, strongly-adsorbed layer. The architecture of this layer was studied by neutron reflection. Data on dried adsorbed layers indicates a dense PMMA film whose thickness gradually increases with annealing time in the melt from a minimal value. Evidently, annealing gradually relaxes a rather flat non-equilibrium structure produced by spin-coating. The thicknesses, h, in a series of dry layers annealed long enough to achieve equilibrium conditions in the melt scale as h ∼ N1/2. Data on swollen layers suggest a dilute, extended layer, but the preliminary results cannot give a definitive confirmation of the brush structure predicted by Guiselin.11
APA, Harvard, Vancouver, ISO, and other styles
50

Wang, Dali, Qiwei Wang, Yuyan Wang, Peiru Chen, Xueguang Lu, Fei Jia, Yehui Sun, et al. "Targeting oncogenic KRAS with molecular brush-conjugated antisense oligonucleotides." Proceedings of the National Academy of Sciences 119, no. 29 (July 14, 2022). http://dx.doi.org/10.1073/pnas.2113180119.

Full text
Abstract:
The mutant form of the guanosine triphosphatase (GTPase) KRAS is a key driver in human tumors but remains a challenging therapeutic target, making KRAS MUT cancers a highly unmet clinical need. Here, we report a class of bottlebrush polyethylene glycol (PEG)–conjugated antisense oligonucleotides (ASOs) for potent in vivo KRAS depletion. Owing to their highly branched architecture, these molecular nanoconstructs suppress nearly all side effects associated with DNA–protein interactions and substantially enhance the pharmacological properties of the ASO, such as plasma pharmacokinetics and tumor uptake. Systemic delivery to mice bearing human non–small-cell lung carcinoma xenografts results in a significant reduction in both KRAS levels and tumor growth, and the antitumor performance well exceeds that of current popular ASO paradigms, such as chemically modified oligonucleotides and PEGylation using linear or slightly branched PEG. Importantly, these conjugates relax the requirement on the ASO chemistry, allowing unmodified, natural phosphodiester ASOs to achieve efficacy comparable to that of chemically modified ones. Both the bottlebrush polymer and its ASO conjugates appear to be safe and well tolerated in mice. Together, these data indicate that the molecular brush–ASO conjugate is a promising therapeutic platform for the treatment of KRAS -driven human cancers and warrant further preclinical and clinical development.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography