Phosphonic Acid-Functionalized Diblock Copolymer Nano-Objects via Polymerization-Induced Self-Assembly: Synthesis, Characterization, and Occlusion into Calcite CrystalsClick to copy article linkArticle link copied!
Abstract
Dialkylphosphonate-functionalized and phosphonic acid-functionalized macromolecular chain transfer agents (macro-CTAs) are utilized for the reversible addition–fragmentation chain transfer (RAFT) dispersion polymerization of benzyl methacrylate (BzMA) at 20% w/w solids in methanol at 64 °C. Spherical, worm-like, and vesicular nano-objects could each be generated through systematic variation of the mean degree of polymerization of the core-forming PBzMA block when using relatively short macro-CTAs. Construction of detailed phase diagrams is essential for the reproducible targeting of pure copolymer morphologies, which were characterized using transmission electron microscopy (TEM) and dynamic light scattering (DLS). For nano-objects prepared using the phosphonic acid-based macro-CTA, transfer from methanol to water leads to the development of anionic surface charge as a result of ionization of the stabilizer chains, but this does not adversely affect the copolymer morphology. Given the well-known strong affinity of phosphonic acid for calcium ions, selected nano-objects were evaluated for their in situ occlusion within growing CaCO3 crystals. Scanning electron microscopy (SEM) studies provide compelling evidence for the occlusion of both worm-like and vesicular phosphonic acid-based nano-objects and hence the production of a series of interesting new organic–inorganic nanocomposites.
Introduction
Experimental Section
Materials
Synthesis of Poly(methacryloyloxymethyl dimethylphosphonate) [PMP] Macro-CTA
Hydrolysis of Poly(methacryloyloxymethyl dimethylphosphonate) [PMP] Yielding Poly(methacryloxymethylphosphonic acid) [PMPA]
Chain Extension Experiments with PMPA Macro-CTAs
Diblock Copolymer Synthesis via Alcoholic Dispersion Polymerization
CaCO3 Precipitation via the Ammonia Diffusion Method (79)
Characterization
NMR Spectroscopy
Gel Permeation Chromatography (GPC)
Dynamic Light Scattering (DLS) and Zeta Potential Measurements
Transmission Electron Microscopy (TEM)
Optical Microscopy, Scanning Electron Microscopy (SEM), and Raman Microscopy
Results and Discussion
Synthesis of Poly(methacryloyloxymethyl dimethylphosphonate) and Poly(methacryloyloxymethyl phosphonic acid) Macro-CTAs
Polymerization Induced Self-Assembly (PISA) in Aqueous and Alcoholic Media
GPC | DLS | TEM | |||||
---|---|---|---|---|---|---|---|
target composition | BzMA % conv | BzMA DP | Mn [kg mol–1] | Mw/Mn | Dh [nm] | PDI | morphology |
PMP24–PBzMA50 | 99 | 50 | 9.8 | 1.17 | 24 | 0.29 | S |
PMP24–PBzMA58 | 99 | 57 | 11.0 | 1.15 | 35 | 0.43 | S + W |
PMP24–PBzMA65 | >99 | 65 | 11.7 | 1.15 | 812 | 0.99 | S + W |
PMP24–PBzMA72 | >99 | 72 | 12.4 | 1.15 | 312 | 0.55 | S + W + V |
PMP24–PBzMA80 | >99 | 80 | 13.3 | 1.14 | 238 | 0.30 | W + V |
PMP24–PBzMA100 | >99 | 100 | 15.8 | 1.14 | 352 | 0.12 | V |
PMP24–PBzMA200 | >99 | 200 | 28.1 | 1.11 | 201 | 0.20 | V |
PMP24–PBzMA300 | >99 | 300 | 34.7 | 1.21 | 181 | 0.07 | V |
PMP32–PBzMA50 | 96 | 48 | 10.6 | 1.19 | 17 | 0.09 | S |
PMP32–PBzMA65 | 94 | 61 | 12.0 | 1.20 | 29 | 0.26 | S + W |
PMP32–PBzMA70 | 97 | 68 | 12.6 | 1.17 | 84 | 0.28 | S + W |
PMP32–PBzMA80 | 99 | 79 | 13.7 | 1.18 | 249 | 0.31 | W |
PMP32–PBzMA90 | 99 | 89 | 14.4 | 1.19 | 499 | 0.47 | W + V |
PMP32–PBzMA100 | >99 | 100 | 16.0 | 1.15 | 356 | 0.44 | W + V |
PMP32–PBzMA120 | >99 | 120 | 19.3 | 1.16 | 167 | 0.09 | W + V |
PMP32–PBzMA135 | >99 | 135 | 20.6 | 1.18 | 175 | 0.17 | W + V |
PMP32–PBzMA150 | 97 | 146 | 22.5 | 1.15 | 158 | 0.07 | V |
PMP32–PBzMA200 | >99 | 200 | 27.7 | 1.16 | 152 | 0.03 | V |
PMP32–PBzMA300 | >99 | 300 | 39.2 | 1.20 | 136 | 0.03 | V |
PMP42–PBzMA50 | >99 | 50 | 14.7 | 1.18 | 16 | 0.16 | S |
PMP42–PBzMA80 | >99 | 80 | 22.7 | 1.17 | 23 | 0.12 | S |
PMP42–PBzMA100 | >99 | 100 | 25.9 | 1.18 | 27 | 0.13 | S |
PMP42–PBzMA200 | 98 | 196 | 44.1 | 1.17 | 53 | 0.08 | S |
PMP42–PBzMA300 | >99 | 300 | 63.1 | 1.19 | 63 | 0.11 | S |
PMP42–PBzMA600 | 93 | 558 | 109.5 | 1.53 | 76 | 0.06 | S |
Generation of Organic–Inorganic Hybrid Materials
DLS | TEM | ||||
---|---|---|---|---|---|
target composition | BzMA % conv | BzMA DP | Dh [nm] | PDI | morphology |
PMPA24–PBzMA35 | >99 | 35 | 15 | 0.13 | S |
PMPA24–PBzMA45 | 96 | 43 | 61 | 0.29 | S + W |
PMPA24–PBzMA50 | >99 | 50 | 421 | 0.61 | S + W + V |
PMPA24–PBzMA60 | 94 | 56 | 556 | 0.29 | S + L + V |
PMPA24–PBzMA70 | >99 | 70 | 429 | 0.91 | S + L + V |
PMPA24–PBzMA100 | >99 | 100 | 459 | 0.18 | V |
PMPA24–PBzMA200 | >99 | 200 | 439 | 0.14 | V |
PMPA24–PBzMA300 | >99 | 300 | 529 | 0.06 | V |
PMPA32–PBzMA35 | >99 | 35 | 17 | 0.10 | S |
PMPA32–PBzMA42 | 96 | 40 | 38 | 0.23 | S + W |
PMPA32–PBzMA50 | >99 | 50 | 227 | 0.57 | S + W |
PMPA32–PBzMA56 | >99 | 56 | 517 | 0.68 | W + L + V |
PMPA32–PBzMA70 | 91 | 64 | 405 | 0.16 | W + V |
PMPA32–PBzMA72 | >99 | 72 | 355 | 0.24 | W + V |
PMPA32–PBzMA100 | 98 | 98 | 309 | 0.23 | V |
PMPA32–PBzMA200 | >99 | 200 | 282 | 0.14 | V |
PMPA32–PBzMA300 | >99 | 300 | 212 | 0.07 | V |
PMPA42–PBzMA35 | 98 | 34 | 21 | 0.08 | S |
PMPA42–PBzMA50 | 90 | 45 | 174 | 0.29 | W |
PMPA42–PBzMA55 | >99 | 55 | 267 | 0.29 | W |
PMPA42–PBzMA70 | 93 | 65 | 248 | 0.22 | W + V |
PMPA42–PBzMA72 | >99 | 72 | 246 | 0.21 | W + V |
PMPA42–PBzMA100 | 93 | 93 | 194 | 0.16 | V |
PMPA42–PBzMA200 | >99 | 200 | 264 | 0.10 | V |
PMPA42–PBzMA300 | >99 | 300 | 214 | 0.02 | V |
Conclusions
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.5b02212.
31P NMR spectra and GPC data for the three PMPA macro-CTAs, GPC and DLS data for kinetics of dispersion polymerization experiments, GPC data for all PMPx–PBzMAy diblock copolymers, additional TEM images, DLS data for aqueous dispersion polymerizations conducted using binary mixtures of macro-CTAs, and Raman spectra for selected CaCO3 occlusion experiments (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
EPSRC is acknowledged for postdoctoral support of A.H. (EP/K006290/1) and A.N.K. (EP/K006304/1). The authors thank André H. Gröschel for the preparation of the nano-object illustrations.
Added in Proof
The following reference has been added by the author. Reference 98 was inserted at galley stage. Y. Ning, L. A. Fielding, T. S. Andrews, D. J. Growney and S. P. Armes Sulfate-based anionic diblock copolymer nanoparticles for efficient occlusion within zinc oxide. Nanoscale, 2015, 7, 6691−6702.
References
This article references 98 other publications.
- 1Lupa, L.; Popa, A.; Ilia, G. In Phosphorus-Based Polymers: From Synthesis to Applications; The Royal Society of Chemistry: 2014; pp 225– 251.Google ScholarThere is no corresponding record for this reference.
- 2Clearfield, A. Curr. Opin. Solid State Mater. Sci. 2002, 6, 495– 506 DOI: 10.1016/S1359-0286(02)00151-1Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXktVKgtLc%253D&md5=6d83bab221755d65704b8d80dfbdb213Recent advances in metal phosphonate chemistry. IIClearfield, AbrahamCurrent Opinion in Solid State & Materials Science (2003), 6 (6), 495-506CODEN: COSSFX; ISSN:1359-0286. (Elsevier Science Ltd.)A review. Phosphonic acids are more acidic than carboxylic acids. They are nonplanar and possess twice the no. of protons. These attributes lead to strong H bond formation that can be used in design of crystal-engineered structures and self-assembled supramol. arrays. Their superior cation complexing ability gives a vast no. of metal complexes and their use in lanthanide and actinide sepns. Recent work, largely drawn from the work of the author's research group, is summarized in this update of the author's 1996 review [Curr Opin Solid State Mater Sci 1(1996) 268].
- 3van den Brand, J.; Van Gils, S.; Beentjes, P. C. J.; Terryn, H.; Sivel, V.; de Wit, J. H. W. Prog. Org. Coat. 2004, 51, 339– 350 DOI: 10.1016/j.porgcoat.2004.08.005Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVaitLfO&md5=d3624facd5be90053eb453a201fddac0Improving the adhesion between epoxy coatings and aluminum substratesvan den Brand, J.; Van Gils, S.; Beentjes, P. C. J.; Terryn, H.; Sivel, V.; de Wit, J. H. W.Progress in Organic Coatings (2004), 51 (4), 339-350CODEN: POGCAT; ISSN:0300-9440. (Elsevier B.V.)Two methods were developed to improve adhesion over the life of a commonly used epoxy coating on an aluminum substrate. The first method involves application of a 10 nm polymeric layer on the aluminum substrate prior to application of the epoxy coating. The functional groups of the polymer undergo chemisorption onto the oxide surface and are involved in curing of the epoxy resin. The polymers are poly(acrylic acid) (PAA), poly(ethylene-alt-maleic anhydride) (PEMah) and poly(vinyl phosphonic acid) (PvPA). The interface of the epoxy coating and aluminum showed that the polymeric layer was indeed part of the cured product. For poly(ethylene-alt-maleic anhydride), there was a clear region between coating and substrate while for the two other polymers, a weakly cured interface region was formed. Adhesion and durability were also improved by hydration of the aluminum substrate by immersion in boiling water, resulting in formation of a porous pseudo-boehmite oxyhydroxide layer. The epoxy coating fully penetrated into this layer. The adhesion of the epoxy coatings was tested initially and after exposure to 40° water and 40° 5% aq. acetic acid. The poly(ethylene-alt-maleic anhydride)-based system showed good initial adhesion and durability on exposure to water, while the systems based on the other two polymers did not. The pseudoboehmite-based system also showed good initial adhesion and durability in the presence of water. None of the proposed systems were able to withstand 40° 5% acetic acid and severe corrosion of Al occurred underneath the epoxy coating.
- 4Zhang, Q. M.; Serpe, M. J. Macromolecules 2014, 47, 8018– 8025 DOI: 10.1021/ma501997xGoogle ScholarThere is no corresponding record for this reference.
- 5Alexandrino, E. M.; Ritz, S.; Marsico, F.; Baier, G.; Mailander, V.; Landfester, K.; Wurm, F. R. J. Mater. Chem. B 2014, 2, 1298– 1306 DOI: 10.1039/c3tb21295eGoogle Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXisFSitro%253D&md5=1b741a14fff5f826b14b79d89053ca76Paclitaxel-loaded polyphosphate nanoparticles: a potential strategy for bone cancer treatmentAlexandrino, Evandro M.; Ritz, Sandra; Marsico, Filippo; Baier, Grit; Mailaender, Volker; Landfester, Katharina; Wurm, Frederik R.Journal of Materials Chemistry B: Materials for Biology and Medicine (2014), 2 (10), 1298-1306CODEN: JMCBDV; ISSN:2050-7518. (Royal Society of Chemistry)While it has been shown that phosphates can target mols. and nanocarriers to bone we herein demonstrate the prepn. of polyphosphate nanoparticles loaded with paclitaxel using a simple miniemulsion/solvent-evapn. technique as a model for chemotherapeutic delivery. Polyphosphates exhibit much higher structural versatility, relying on the pentavalence of the phosphorus center compared to conventional polyesters. This versatility allows for the development of new degradable polymeric carriers with inherent bone adhesion ability by the interaction of the nanoparticles with a calcium phosphate material used for bone regeneration. The novel polyphosphate nanoparticles were investigated in detail with respect to their size distribution, zeta-potential, thermal and morphol. properties and were further proven to be efficiently loaded with a hydrophobic drug (up to 15 wt%). The in vitro cytotoxicity was assessed against human cancer cell lines (HeLa and Saos-2), and the paclitaxel-loaded nanoparticles showed a similar cytotoxicity profile similar to the com. available formulation Taxomedac and the pure paclitaxel for loading ratios of 10 wt% but addnl. proved efficient adhesion on calcium phosphate granules allowing drug delivery to bone. This first report demonstrates that polyphosphate nanoparticles are promising materials for the development of systemic or local bone cancer treatment, even by direct application or by formation of composites with calcium phosphate cements.
- 6Brunklaus, G.; Schauff, S.; Markova, D.; Klapper, M.; Müllen, K.; Spiess, H.-W. J. Phys. Chem. B 2009, 113, 6674– 6681 DOI: 10.1021/jp901714fGoogle ScholarThere is no corresponding record for this reference.
- 7Quittmann, U.; Lecamp, L.; El Khatib, W.; Youssef, B.; Bunel, C. Macromol. Chem. Phys. 2001, 202, 628– 635 DOI: 10.1002/1521-3935(20010301)202:5<628::AID-MACP628>3.0.CO;2-TGoogle ScholarThere is no corresponding record for this reference.
- 8Canadell, J.; Hunt, B. J.; Cook, A. G.; Mantecón, A.; Cádiz, V. Polym. Degrad. Stab. 2007, 92, 1482– 1490 DOI: 10.1016/j.polymdegradstab.2007.05.017Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXos1Sis70%253D&md5=42925672d0dbcdd3261acdd9d2d44200Flame retardance and shrinkage reduction of polystyrene modified with acrylate-containing phosphorus and crosslinkable spiro-orthoester moietiesCanadell, J.; Hunt, B. J.; Cook, A. G.; Mantecon, A.; Cadiz, V.Polymer Degradation and Stability (2007), 92 (8), 1482-1490CODEN: PDSTDW; ISSN:0141-3910. (Elsevier Ltd.)Styrene was radically copolymd. with a spiro-orthoester with an acrylate group (SOE-AC), and terpolymd. with SOE-AC and diethyl(methacryloyloxymethyl)phosphonate (DEMMP). This was done for several different feed ratios, to obtain polymers with spiro-orthoester moieties in the side chain. These polymers were then crosslinked with ytterbium triflate, as cationic initiator, via the double ring-opening polymn. The thermal stability and fire retardant properties of these materials were evaluated by TGA and LOI. The DEMMP-contg. polymers give materials which were significantly more flame retardant than the nonphosphorus-contg. materials, as indicated by the LOI measurements. The vol. changes measured upon crosslinking of the polymers were evaluated by d. measurements with a gas pycnometer. In all the cases, expansion was obsd. This indicates that SOE-AC is an effective monomer for crosslinkable polymers without vol. changes.
- 9Köppl, T.; Brehme, S.; Pospiech, D.; Fischer, O.; Wolff-Fabris, F.; Altstädt, V.; Schartel, B.; Döring, M. J. Appl. Polym. Sci. 2013, 128, 3315– 3324 DOI: 10.1002/app.38520Google ScholarThere is no corresponding record for this reference.
- 10Monge, S.; Canniccioni, B.; Graillot, A.; Robin, J.-J. Biomacromolecules 2011, 12, 1973– 1982 DOI: 10.1021/bm2004803Google ScholarThere is no corresponding record for this reference.
- 11Lin, Z.; Cao, S.; Chen, X.; Wu, W.; Li, J. Biomacromolecules 2013, 14, 2206– 2214 DOI: 10.1021/bm4003442Google ScholarThere is no corresponding record for this reference.
- 12David, G.; Asri, Z. E.; Rich, S.; Castignolles, P.; Guillaneuf, Y.; Lacroix-Desmazes, P.; Boutevin, B. Macromol. Chem. Phys. 2009, 210, 631– 639 DOI: 10.1002/macp.200800540Google ScholarThere is no corresponding record for this reference.
- 13Canniccioni, B.; Monge, S.; David, G.; Robin, J.-J. Polym. Chem. 2013, 4, 3676– 3685 DOI: 10.1039/c3py00426kGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXovFWktb0%253D&md5=ca49c191735fec9f1aac492dc62b29b2RAFT polymerization of dimethyl(methacryloyloxy)methyl phosphonate and its phosphonic acid derivative: a new opportunity for phosphorus-based materialsCanniccioni, Benjamin; Monge, Sophie; David, Ghislain; Robin, Jean-JacquesPolymer Chemistry (2013), 4 (13), 3676-3685CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)The RAFT polymn. of dimethyl(methacryloyloxy)methyl phosphonate (MAPC1) using dithioester chain transfer agents is reported for the first time to our knowledge in the literature. Poly(dimethyl(methacryloyloxy)methyl phosphonate) (PMAPC1) was synthesized in DMF at 70° with a good control over the mol. wt., the latter ranging from 8000 to 24 000 g mol-1. Polymers were characterized by 1H and 31P NMR, and size exclusion chromatog. using both refractive index and triple detection. The effect of the solvent was also investigated. We demonstrated that polarity had an effect on the control of the polymn. as low polarity led to low polymn. rate and termination reactions whereas very high polarity resulted in high polymn. rate but also transfer reactions. Addnl., we showed that it was possible to restart the RAFT polymn. from a PMAPC1 macrochain transfer agent. PMAPC1 was hydrolyzed to afford poly((methacryloyloxy)methyl phosphonic acid) (hPMAPC1), whose pKa values were detd. Finally, we achieved the controlled RAFT polymn. of the (methacryloyloxy)methyl phosphonic acid (hMAPC1). The easy obtaining of PMAPC1 and hPMAPC1 opens the way to the synthesis of complex polymer architectures.
- 14Graillot, A.; Monge, S.; Faur, C.; Bouyer, D.; Robin, J.-J. Polym. Chem. 2013, 4, 795– 803 DOI: 10.1039/C2PY20720FGoogle Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtl2nsQ%253D%253D&md5=4bd338a8e77928e0045f58ec2816a9edSynthesis by RAFT of innovative well-defined (co)polymers from a novel phosphorus-based acrylamide monomerGraillot, Alain; Monge, Sophie; Faur, Catherine; Bouyer, Denis; Robin, Jean-JacquesPolymer Chemistry (2013), 4 (3), 795-803CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)The present contribution reports on the synthesis and controlled polymn. of a novel acrylamide monomer contg. phosphonated moieties, namely diethyl-2-(acrylamido)ethylphosphonate. This monomer appears to be of great interest due to the phosphonated moieties, which can lead to a wide range of applications, assocd. with the chem. stability of the acrylamide compared to more common (meth)acrylate monomers. Reversible Addn.-Fragmentation Transfer (RAFT) polymn. of this monomer was studied using two different trithiocarbonate chain transfer agents, and it allowed the synthesis of poly(diethyl-2-(acrylamido)ethylphosphonate) with controlled mol. wt. and low dispersity. Addnl., a diblock copolymer was successfully prepd. by a similar RAFT procedure using thermosensitive poly(N-n-propylacrylamide) as a macro-chain transfer agent. A combination of both stimuli-responsive and phosphonated ester or phosphonic diacid (after hydrolysis) contg. blocks appears valuable for drug delivery or water treatment, for instance.
- 15Grøndahl, L.; Suzuki, S.; Wentrup-Byrne, E. Chem. Commun. 2008, 3314– 3316 DOI: 10.1039/b803223hGoogle ScholarThere is no corresponding record for this reference.
- 16Miyatake, K.; Hay, A. S. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 1854– 1859 DOI: 10.1002/pola.1163Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXjsVaitLo%253D&md5=d4141b680795dc56cc6be9b3fe953aecSynthesis of novel phosphinic acid-containing polymersMiyatake, Kenji; Hay, Allan S.Journal of Polymer Science, Part A: Polymer Chemistry (2001), 39 (11), 1854-1859CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)Three arylene difluoride monomers contg. phosphine oxide (1), phosphinic acid (2), or phosphinate ester (3) groups were prepd. and polymd. with bisphenol A to give novel poly(arylene ether)s (4, 5, and 6). The polymers obtained had moderate mol. wts. (ηinh: 0.14-0.30 dL g-1 in N-methylpyrrolidinone) and glass-transition temps. (Tg: 102-200°), depending on the phosphine group in the main chain. Using bis(4-fluorophenyl)sulfone as a comonomer improved the polymn. to give copolymers with higher soln. viscosities. The stoichiometric investigation revealed that 7 mol % excess of fluoride monomer gave the highest mol. wt. copolymer 8 with ηinh of 0.78 dL g-1, which had a Tg of 176 °C, a Td10% of 432°, and formed a hard film by casting from soln.
- 17Watson, J. D.; Crick, F. H. C. Nature 1974, 248, 765– 765 DOI: 10.1038/248765a0Google ScholarThere is no corresponding record for this reference.
- 18Ornelas-Megiatto, C.; Wich, P. R.; Fréchet, J. M. J. J. Am. Chem. Soc. 2012, 134, 1902– 1905 DOI: 10.1021/ja207366kGoogle ScholarThere is no corresponding record for this reference.
- 19Uchiyama, T.; Kiritoshi, Y.; Watanabe, J.; Ishihara, K. Biomaterials 2003, 24, 5183– 5190 DOI: 10.1016/S0142-9612(03)00441-1Google ScholarThere is no corresponding record for this reference.
- 20Pegoraro, C.; Cecchin, D.; Gracia, L. S.; Warren, N.; Madsen, J.; Armes, S. P.; Lewis, A.; MacNeil, S.; Battaglia, G. Cancer Lett. 2013, 334, 328– 337 DOI: 10.1016/j.canlet.2013.02.007Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXktlKksLk%253D&md5=b83b822803e34fa092adba772af91f37Enhanced drug delivery to melanoma cells using PMPC-PDPA polymersomesPegoraro, Carla; Cecchin, Denis; Gracia, Lorena Simon; Warren, Nicholas; Madsen, Jeppe; Armes, Steven P.; Lewis, Andrew; MacNeil, Sheila; Battaglia, GiuseppeCancer Letters (New York, NY, United States) (2013), 334 (2), 328-337CODEN: CALEDQ; ISSN:0304-3835. (Elsevier)We present the efficient and stable encapsulation of doxorubicin within pH sensitive polymeric vesicles (polymersomes) for intracellular and nuclear delivery to melanoma cells. We demonstrate that PMPC25-PDPA70 polymersomes can encapsulate doxorubicin for long periods of time without significant drug release. We demonstrate that empty polymersomes are non-toxic and that they are quickly and more efficiently internalised by melanoma cells compared to healthy cells. Encapsulated doxorubicin has a strong cytotoxic effect on both healthy and cancerous cells, but when encapsulated it had a preferential effect on melanoma cells indicating that this formulation can be used to achieve an enhanced drug delivery to cancerous cells rather than to the healthy surrounding cells.
- 21Hemp, S. T.; Smith, A. E.; Bryson, J. M.; Allen, M. H.; Long, T. E. Biomacromolecules 2012, 13, 2439– 2445 DOI: 10.1021/bm300689fGoogle ScholarThere is no corresponding record for this reference.
- 22Liu, G.; Jin, Q.; Liu, X.; Lv, L.; Chen, C.; Ji, J. Soft Matter 2011, 7, 662– 669 DOI: 10.1039/C0SM00708KGoogle ScholarThere is no corresponding record for this reference.
- 23Matyjaszewski, K. Macromolecules 2012, 45, 4015– 4039 DOI: 10.1021/ma3001719Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XlsVaqs7w%253D&md5=350b580bd1bb46c21ba5dbfd65a6811dAtom Transfer Radical Polymerization (ATRP): Current Status and Future PerspectivesMatyjaszewski, KrzysztofMacromolecules (Washington, DC, United States) (2012), 45 (10), 4015-4039CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A review. Current status and future perspectives in atom transfer radical polymn. (ATRP) are presented. Special emphasis is placed on mechanistic understanding of ATRP, recent synthetic and process development, and new controlled polymer architectures enabled by ATRP. New hybrid materials based on org./inorg. systems and natural/synthetic polymers are presented. Some current and forthcoming applications are described.
- 24Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2012, 65, 985– 1076 DOI: 10.1071/CH12295Google ScholarThere is no corresponding record for this reference.
- 25Lobb, E. J.; Ma, I.; Billingham, N. C.; Armes, S. P.; Lewis, A. L. J. Am. Chem. Soc. 2001, 123, 7913– 7914 DOI: 10.1021/ja003906dGoogle ScholarThere is no corresponding record for this reference.
- 26Madsen, J.; Canton, I.; Warren, N. J.; Themistou, E.; Blanazs, A.; Ustbas, B.; Tian, X.; Pearson, R.; Battaglia, G.; Lewis, A. L.; Armes, S. P. J. Am. Chem. Soc. 2013, 135, 14863– 14870 DOI: 10.1021/ja407380tGoogle ScholarThere is no corresponding record for this reference.
- 27Yusa, S.-i.; Fukuda, K.; Yamamoto, T.; Ishihara, K.; Morishima, Y. Biomacromolecules 2005, 6, 663– 670 DOI: 10.1021/bm0495553Google ScholarThere is no corresponding record for this reference.
- 28Ohno, S.; Hasegawa, S.; Liu, H.; Ishihara, K.; Yusa, S.-i. Polym. J. 2015, 47, 71 DOI: 10.1038/pj.2014.92Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVGntQ%253D%253D&md5=4b23620ea73ecacb0ab49ffed37613e5Aggregation behavior in water of amphiphilic diblock copolymers bearing biocompatible phosphorylcholine and cholesteryl groupsOhno, Sayaka; Hasegawa, Shoto; Liu, Huihua; Ishihara, Kazuhiko; Yusa, Shin-ichiPolymer Journal (Tokyo, Japan) (2015), 47 (1), 71-76CODEN: POLJB8; ISSN:0032-3896. (NPG Nature Asia-Pacific)Poly(2-(methacryloyloxy)ethyl phosphorylcholine)-block-poly(cholesteryl 6-methacryloyloxyhexanoate) (PMPC82-b-PChMn) copolymers with different PChM block lengths were prepd. via reversible addn.-fragmentation chain transfer controlled/living radical polymn. using a PMPC-based macro-chain transfer agent. The subscript no. and n (=3 and 6) refer to the d.p. of the PMPC and PChM blocks, resp. PMPC82-b-PChMn cannot dissolve in water directly due to the strong hydrophobic nature of the PChM block. To prep. the aq. soln., the diblock copolymer was dissolved in an org. solvent and then dialyzed against pure water. These diblock copolymers formed spherical and rod-like micelles in water, depending on the compn. of cholesteryl (Chol) group in the polymer. The prepd. aggregates were characterized using static light scattering, dynamic light scattering, transmission electron microscopy and fluorescence probe techniques. The characterization results suggest that the morphol. of the polymer aggregates can be controlled from spherical to rod-like micelles by increasing the no. of Chol groups in the polymer.
- 29Huang, J.; Matyjaszewski, K. Macromolecules 2005, 38, 3577– 3583 DOI: 10.1021/ma047564yGoogle ScholarThere is no corresponding record for this reference.
- 30Suzuki, S.; Whittaker, M. R.; Grøndahl, L.; Monteiro, M. J.; Wentrup-Byrne, E. Biomacromolecules 2006, 7, 3178– 3187 DOI: 10.1021/bm060583qGoogle ScholarThere is no corresponding record for this reference.
- 31Blidi, I.; Geagea, R.; Coutelier, O.; Mazieres, S.; Violleau, F.; Destarac, M. Polym. Chem. 2012, 3, 609– 612 DOI: 10.1039/c2py00541gGoogle Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvFalurc%253D&md5=cbc2b11df46ed53d703cf1ef26916078Aqueous RAFT/MADIX polymerization of vinylphosphonic acidBlidi, Issam; Geagea, Roland; Coutelier, Olivier; Mazieres, Stephane; Violleau, Frederic; Destarac, MathiasPolymer Chemistry (2012), 3 (3), 609-612CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)RAFT/MADIX polymn. of vinylphosphonic acid (VPA) was controlled in water with an O-Et xanthate transfer agent. This represents the first example of reversible deactivation radical polymn. of a monomer bearing an unprotected phosphonic acid function. Hence, macromol. engineering of polyphosphonates can now be envisioned by directly polymg. VPA in water.
- 32Markova, D.; Kumar, A.; Klapper, M.; Müllen, K. Polymer 2009, 50, 3411– 3421 DOI: 10.1016/j.polymer.2009.06.011Google ScholarThere is no corresponding record for this reference.
- 33David, G.; Negrell, C.; Manseri, A.; Boutevin, B. J. Appl. Polym. Sci. 2009, 114, 2213– 2220 DOI: 10.1002/app.30438Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVCgtb3K&md5=a82c0421fd3862209b38be8f247f8c41Poly(MMA)-b-poly(monophosphonic acrylate) diblock copolymers obtained by ATRP and used as additives for anticorrosive coatingsDavid, Ghislain; Negrell, Claire; Manseri, Abdelatif; Boutevin, BernardJournal of Applied Polymer Science (2009), 114 (4), 2213-2220CODEN: JAPNAB; ISSN:0021-8995. (John Wiley & Sons, Inc.)Atom transfer radical polymn. (ATRP) of dimethyl(methacryloyloxy)methyl phosphonate (MAC1P) was investigated in toluene, in the presence of Me 2-bromoisobutyrate as the initiator, and using different metal and ligand systems. Polymn. proceeded with very low monomer conversion, which was attributed to the ability of phosphorus to complex the copper ions, removing copper ions from original ligand, and then stopping the MAC1P polymn. Poly(MMA)-b-poly(phosphonate acrylate) diblock copolymer structure was efficiently obtained by the ATRP process, based on a four-step reaction. Poly(MMA)-b-poly(tert-Bu acrylate) diblock copolymer was first obtained by ATRP, then the tert-Bu groups were removed and phosphonate functions were incorporated by esterification reaction, using 4-dimethylaminopyridine as the catalyst. This new diblock copolymer was used as an additive for anticorrosive coating; however, no improvement (using the salt spray test technique) was obsd. comparatively with the statistical copolymer with the same acid content. P-contg. additive led to the increased corrosion resistance. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009.
- 34Mukumoto, K.; Zhong, M.; Matyjaszewski, K. Eur. Polym. J. 2014, 56, 11– 16 DOI: 10.1016/j.eurpolymj.2014.03.029Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpslCrsrc%253D&md5=38541cf2544ad1e721a9c69fe4c33cb9Atom transfer radical polymerization of dimethyl(methacryloyloxymethyl) phosphonateMukumoto, Kosuke; Zhong, Mingjiang; Matyjaszewski, KrzysztofEuropean Polymer Journal (2014), 56 (), 11-16CODEN: EUPJAG; ISSN:0014-3057. (Elsevier Ltd.)Atom transfer radical polymn. (ATRP) of dimethyl(methacryloyloxymethyl) phosphonate (DMMAMP, also referred to as MAPC1) with a copper/2,2'-bipyridine based catalytic system in DMSO resulted in a homopolymer with narrow mol. wt. distribution. Initiators for continuous activator regeneration (ICAR) ATRP of this monomer was also successful when using tris[(2-pyridyl)methyl]amine as ligand and azobisisobutyronitrile as supplementary initiator for the regeneration of the CuI activator. Normal ATRP provided better control compared to ICAR ATRP. Differential scanning calorimetry measurements of the resulting poly(DMMAMP) revealed a glass transition temp. (Tg = 63 °C), lower than that of poly(Me methacrylate). Chain extension of the poly(dimethyl(methacryloyloxymethyl) phosphonate) macroinitiator with styrene was successfully achieved with a high fraction of block copolymer.
- 35Charleux, B.; Delaittre, G.; Rieger, J.; D’Agosto, F. Macromolecules 2012, 45, 6753– 6765 DOI: 10.1021/ma300713fGoogle ScholarThere is no corresponding record for this reference.
- 36Warren, N. J.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 10174– 10185 DOI: 10.1021/ja502843fGoogle Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVGlu7bM&md5=a8ab4eb77ba9f218d45dc9cbb4e9fbf7Polymerization-Induced Self-Assembly of Block Copolymer Nano-objects via RAFT Aqueous Dispersion PolymerizationWarren, Nicholas J.; Armes, Steven P.Journal of the American Chemical Society (2014), 136 (29), 10174-10185CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A review. In this Perspective, we discuss the recent development of polymn.-induced self-assembly mediated by reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. This approach has quickly become a powerful and versatile technique for the synthesis of a wide range of bespoke org. diblock copolymer nano-objects of controllable size, morphol., and surface functionality. Given its potential scalability, such environmentally-friendly formulations are expected to offer many potential applications, such as novel Pickering emulsifiers, efficient microencapsulation vehicles, and sterilizable thermo-responsive hydrogels for the cost-effective long-term storage of mammalian cells.
- 37Sun, J.-T.; Hong, C.-Y.; Pan, C.-Y. Polym. Chem. 2013, 4, 873– 881 DOI: 10.1039/C2PY20612AGoogle Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlelt7g%253D&md5=6d4686d05496bb207018c7ec714735fdRecent advances in RAFT dispersion polymerization for preparation of block copolymer aggregatesSun, Jiao-Tong; Hong, Chun-Yan; Pan, Cai-YuanPolymer Chemistry (2013), 4 (4), 873-881CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A review. Differently from bulk, soln., suspension, emulsion, and miniemulsion polymns., the controlled radical dispersion polymn. (CRDP) demonstrates self-assembly of the block copolymers formed in the homogeneous system, forming various kinds of micelles or vesicles. Thus, this technol. can prep. both the block copolymers and the polymeric aggregates directly. Among CRDP, the reversible addn.-fragmentation chain transfer (RAFT) dispersion polymn. has been studied in relative detail and has been successfully developed to prep. a diverse range of assemblies. Several typical systems for RAFT dispersion polymn. are presented in detail and the factors influencing the polymn. and the in situ self-assembly are also highlighted in this minireview.
- 38Cunningham, V. J.; Alswieleh, A. M.; Thompson, K. L.; Williams, M.; Leggett, G. J.; Armes, S. P.; Musa, O. M. Macromolecules 2014, 47, 5613– 5623 DOI: 10.1021/ma501140hGoogle ScholarThere is no corresponding record for this reference.
- 39Blanazs, A.; Ryan, A. J.; Armes, S. P. Macromolecules 2012, 45, 5099– 5107 DOI: 10.1021/ma301059rGoogle ScholarThere is no corresponding record for this reference.
- 40An, Z.; Shi, Q.; Tang, W.; Tsung, C.-K.; Hawker, C. J.; Stucky, G. D. J. Am. Chem. Soc. 2007, 129, 14493– 14499 DOI: 10.1021/ja0756974Google ScholarThere is no corresponding record for this reference.
- 41Rieger, J.; Grazon, C.; Charleux, B.; Alaimo, D.; Jérôme, C. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 2373– 2390 DOI: 10.1002/pola.23329Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXltVOjsb4%253D&md5=216bf726a4cfe91f8e0935c65fd8d625Pegylated thermally responsive block copolymer micelles and nanogels via in situ RAFT aqueous dispersion polymerizationRieger, Jutta; Grazon, Chloe; Charleux, Bernadette; Alaimo, David; Jerome, ChristineJournal of Polymer Science, Part A: Polymer Chemistry (2009), 47 (9), 2373-2390CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A straightforward approach was developed to synthesize pegylated thermo-responsive core-shell nanoparticles in a min. of steps, directly in water. It is based on RAFT-controlled radical crosslinking copolymn. of N,N-diethylacrylamide (DEAAm) and N,N'-methylene bisacrylamide (MBA) in aq. dispersion polymn. Because DEAAm is water-sol. and poly(N,N-diethylacrylamide) (PDEAAm) exhibits a lower crit. soln. temp. at 32°, the initial medium was homogeneous, whereas the polymer formed a sep. phase at the reaction temp. The first macroRAFT agent was a surface-active trithiocarbonate based on a hydrophilic poly(ethylene oxide) block and a hydrophobic dodecyl chain. It was further extended with N,N-dimethylacrylamide (DMAAm) to target macroRAFT agents with increasing chain length. All macroRAFT agents provided excellent control over the aq. dispersion homopolymn. of DEAAm. When they were used in the radical crosslinking copolymn. of DEAAm and MBA, the stability and size of the resulting gel particles were found to depend strongly on the chain length of the macroRAFT agent, on the concns. of both the monomer and the crosslinker, and on the process (one step or two steps). The best-suited exptl. conditions to reach thermosensitive hydrogels with nanometric size and well-defined surface properties were detd.
- 42Li, Y.; Armes, S. P. Angew. Chem., Int. Ed. 2010, 49, 4042– 4046 DOI: 10.1002/anie.201001461Google ScholarThere is no corresponding record for this reference.
- 43Sugihara, S.; Blanazs, A.; Armes, S. P.; Ryan, A. J.; Lewis, A. L. J. Am. Chem. Soc. 2011, 133, 15707– 15713 DOI: 10.1021/ja205887vGoogle ScholarThere is no corresponding record for this reference.
- 44Liu, G.; Qiu, Q.; Shen, W.; An, Z. Macromolecules 2011, 44, 5237– 5245 DOI: 10.1021/ma200984hGoogle ScholarThere is no corresponding record for this reference.
- 45Chambon, P.; Blanazs, A.; Battaglia, G.; Armes, S. P. Macromolecules 2012, 45, 5081– 5090 DOI: 10.1021/ma300816mGoogle ScholarThere is no corresponding record for this reference.
- 46Ladmiral, V.; Semsarilar, M.; Canton, I.; Armes, S. P. J. Am. Chem. Soc. 2013, 135, 13574– 13581 DOI: 10.1021/ja407033xGoogle ScholarThere is no corresponding record for this reference.
- 47Figg, C. A.; Simula, A.; Gebre, K. A.; Tucker, B. S.; Haddleton, D.; Sumerlin, B. S. Chem. Sci. 2015, 6, 1230 DOI: 10.1039/C4SC03334EGoogle Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvFOitLzI&md5=ddf49e8d73521f3e1d95d304f275f33cPolymerization-induced thermal self-assembly (PITSA)Figg, C. Adrian; Simula, Alexandre; Gebre, Kalkidan A.; Tucker, Bryan S.; Haddleton, David M.; Sumerlin, Brent S.Chemical Science (2015), 6 (2), 1230-1236CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Polymn.-induced self-assembly (PISA) is a versatile technique to achieve a wide range of polymeric nanoparticle morphologies. Most previous examples of self-assembled soft nanoparticle synthesis by PISA rely on a growing solvophobic polymer block that leads to changes in nanoparticle architecture during polymn. in a selective solvent. However, synthesis of block copolymers with a growing stimuli-responsive block to form various nanoparticle shapes has yet to be reported. This new concept using thermo-responsive polymers is termed polymn.-induced thermal self-assembly (PITSA). A reversible addn.-fragmentation chain transfer (RAFT) polymn. of N-isopropylacrylamide from a hydrophilic chain transfer agent composed of N,N-dimethylacrylamide and acrylic acid was carried out in water above the known lower crit. soln. temp. (LCST) of poly(N-isopropylacrylamide) (PNIPAm). After reaching a certain chain length, the growing PNIPAm self-assembled, as induced by the LCST, into block copolymer aggregates within which dispersion polymn. continued. To characterize the nanoparticles at ambient temps. without their dissoln., the particles were crosslinked immediately following polymn. at elevated temps. via the reaction of the acid groups with a diamine in the presence of a carbodiimide. Size exclusion chromatog. was used to evaluate the unimer mol. wt. distributions and reaction kinetics. Dynamic light scattering and transmission electron microscopy provided insight into the size and morphologies of the nanoparticles. The resulting block copolymers formed polymeric nanoparticles with a range of morphologies (e.g., micelles, worms, and vesicles), which were a function of the PNIPAm block length.
- 48Wan, W.-M.; Hong, C.-Y.; Pan, C.-Y. Chem. Commun. 2009, 5883– 5885 DOI: 10.1039/b912804bGoogle ScholarThere is no corresponding record for this reference.
- 49Semsarilar, M.; Jones, E. R.; Blanazs, A.; Armes, S. P. Adv. Mater. 2012, 24, 3378– 3382 DOI: 10.1002/adma.201200925Google ScholarThere is no corresponding record for this reference.
- 50Zehm, D.; Ratcliffe, L. P. D.; Armes, S. P. Macromolecules 2013, 46, 128– 139 DOI: 10.1021/ma301459yGoogle ScholarThere is no corresponding record for this reference.
- 51Karagoz, B.; Boyer, C.; Davis, T. P. Macromol. Rapid Commun. 2014, 35, 417– 421 DOI: 10.1002/marc.201300730Google ScholarThere is no corresponding record for this reference.
- 52Pei, Y.; Dharsana, N. C.; van Hensbergen, J. A.; Burford, R. P.; Roth, P. J.; Lowe, A. B. Soft Matter 2014, 10, 5787– 5796 DOI: 10.1039/C4SM00729HGoogle ScholarThere is no corresponding record for this reference.
- 53Zhang, W.-J.; Hong, C.-Y.; Pan, C.-Y. Macromolecules 2014, 47, 1664– 1671 DOI: 10.1021/ma402497yGoogle ScholarThere is no corresponding record for this reference.
- 54Bleach, R.; Karagoz, B.; Prakash, S. M.; Davis, T. P.; Boyer, C. ACS Macro Lett. 2014, 3, 591– 596 DOI: 10.1021/mz500195uGoogle Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpsFaqs78%253D&md5=ae73c889db0575c204c802e467a5a3a0In Situ Formation of Polymer-Gold Composite Nanoparticles with Tunable MorphologiesBleach, Richard; Karagoz, Bunyamin; Prakash, Shyam M.; Davis, Thomas P.; Boyer, CyrilleACS Macro Letters (2014), 3 (7), 591-596CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)A simple and efficient route to gold-polymer nanoparticle composites is described. Our versatile synthetic route exerts facile control over polymer nanoparticle morphol., including micelles, rod-like structures, and vesicles, all easily attainable from a single polymn. taken to different monomer conversions. Specifically, poly[oligo(ethylene glycol) methacrylate]-b-poly(dimethylaminoethyl methacrylate)-b-poly(styrene) (POEGMA-b-PDMAEMA-b-PST) triblock copolymers were synthesized using a polymn. induced self-assembly (PISA) approach. Subsequently, spherical gold nanoparticles (10 nm AuNPs) were formed at the hydrophilic-hydrophobic nexus of the assembled triblock copolymer nanoaggregates by the addn. of chloroauric acid (HAuCl4) followed by in situ redn. using NaBH4. After redn., the cloudy white nanoparticle dispersions turned to a red-purple color. The gold nanoparticles that formed were stabilized by the enveloping polymeric nanostructures, neither pptn. nor agglomeration occurred. We demonstrated that we were able to tune the gold nanoparticle compn. in these polymer-gold composites by varying the concn. of chloroauric acid. Morphol., particle size, mol. wt., AuNP content, and chem. structure of the polymer structures were characterized by transmittance electron microscopy (TEM), dynamic light scattering (DLS), size exclusion chromatog. (SEC), thermal gravimetric anal. (TGA), and 1H NMR. Finally, the formation of the AuNPs occurred without affecting the polymer nanoparticle morphol.
- 55Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Polym. Chem. 2014, 5, 3466– 3475 DOI: 10.1039/c4py00201fGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmtlGgsr8%253D&md5=5cc9158e0615ba6fe5993d8292ed4880Poly(methacrylic acid)-based AB and ABC block copolymer nano-objects prepared via RAFT alcoholic dispersion polymerizationSemsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P.Polymer Chemistry (2014), 5 (10), 3466-3475CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A series of well-defined amphiphilic poly(methacrylic acid)-poly(benzyl methacrylate) (PMAA-PBzMA) diblock copolymers were synthesized via polymn.-induced self-assembly using an alc. dispersion polymn. formulation. Chain growth was mediated via reversible addn.-fragmentation chain transfer polymn. (RAFT) chem. using a trithiocarbonate-based chain transfer agent (CTA) at 70 °C. The poly(methacrylic acid) block was sol. in ethanol and acts as a steric stabilizer for the growing insol. PBzMA chains, resulting in the in situ generation of diblock copolymer nano-objects in the form of spheres, worms or vesicles, depending on the precise reaction conditions. Copolymer morphologies can be covalently stabilized via crosslinking to prevent their dissocn. when transferred into aq. soln., which leads to the formation of highly anionic nano-objects due to ionization of the PMAA stabilizer chains. ABC triblock copolymer nanoparticles can also be prepd. using this approach, where the third block was based on the semi-fluorinated monomer, 2,2,2-trifluoroethyl methacrylate (TFEMA). GPC studies confirm that chain extension was efficient and high TFEMA conversions can be achieved. Microphase sepn. between the mutually incompatible PBzMA and semi-fluorinated PTFEMA core-forming blocks occurs, producing a range of remarkably complex semi-fluorinated triblock copolymer morphologies.
- 56Gonzato, C.; Semsarilar, M.; Jones, E. R.; Li, F.; Krooshof, G. J. P.; Wyman, P.; Mykhaylyk, O. O.; Tuinier, R.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 11100– 11106 DOI: 10.1021/ja505406sGoogle ScholarThere is no corresponding record for this reference.
- 57Houillot, L.; Bui, C.; Farcet, C. l.; Moire, C.; Raust, J.-A.; Pasch, H.; Save, M.; Charleux, B. ACS Appl. Mater. Interfaces 2010, 2, 434– 442 DOI: 10.1021/am900695aGoogle ScholarThere is no corresponding record for this reference.
- 58Deng, Y.; Yang, C.; Yuan, C.; Xu, Y.; Bernard, J.; Dai, L.; Gérard, J.-F. J. Polym. Sci., Part A: Polym. Chem. 2013, 51, 4558– 4564 DOI: 10.1002/pola.26872Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtF2iu7zN&md5=222a788bfc01b859b7cbb6d15ef1f536Hybrid organic-inorganic block copolymer nano-objects from RAFT polymerization-induced self-assemblyDeng, Yuanming; Yang, Cangjie; Yuan, Conghui; Xu, Yiting; Bernard, Julien; Dai, Lizong; Gerard, Jean-FrancoisJournal of Polymer Science, Part A: Polymer Chemistry (2013), 51 (21), 4558-4564CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A simple route to org.-inorg. (O/I) nano-objects with different morphologies through polymn.-induced block copolymer self-assembly is described. The synthetic strategy relies on the chain-extension of polyhedral oligomeric silsesquioxanes (POSS)-contg. macro-CTA (PMAiBuPOSS13 and PMAiBuPOSS19) with styrene at 120 °C in octane, a selective solvent of the POSS-contg. block. The polymn. system was proven to afford a plethora of O/I nano-objects, such as spherical micelles, cylindrical micelles, and vesicles depending on the resp. molar masses of the PMAiBuPOSS and polystyrene (PS) blocks. The cooling procedure was also proven to be a crucial step to generate particles with a unique morphol. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013.
- 59Fielding, L. A.; Derry, M. J.; Ladmiral, V.; Rosselgong, J.; Rodrigues, A. M.; Ratcliffe, L. P. D.; Sugihara, S.; Armes, S. P. Chem. Sci. 2013, 4, 2081– 2087 DOI: 10.1039/c3sc50305dGoogle Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXltVOltrs%253D&md5=03d448f80f5c6e8a2ed9acf3b6fc8abbRAFT dispersion polymerization in non-polar solvents: facile production of block copolymer spheres, worms and vesicles in n-alkanesFielding, Lee A.; Derry, Matthew J.; Ladmiral, Vincent; Rosselgong, Julien; Rodrigues, Aurelie M.; Ratcliffe, Liam P. D.; Sugihara, Shinji; Armes, Steven P.Chemical Science (2013), 4 (5), 2081-2087CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Well-defined poly(lauryl methacrylate-benzyl methacrylate) (PLMA-PBzMA) diblock copolymer nanoparticles are prepd. in n-heptane at 90° via reversible addn.-fragmentation chain transfer (RAFT) polymn. Under these conditions, the PLMA macromol. chain transfer agent (macro-CTA) is sol. in n-heptane, whereas the growing PBzMA block quickly becomes insol. Thus this dispersion polymn. formulation leads to polymn.-induced self-assembly (PISA). Using a relatively long PLMA macro-CTA with a mean d.p. (DP) of 37 or higher leads to the formation of well-defined spherical nanoparticles of 41 to 139 nm diam., depending on the DP targeted for the PBzMA block. In contrast, TEM studies confirm that using a relatively short PLMA macro-CTA (DP = 17) enables both worm-like and vesicular morphologies to be produced, in addn. to the spherical phase. A detailed phase diagram has been elucidated for this more asym. diblock copolymer formulation, which ensures that each pure phase can be targeted reproducibly. 1H NMR spectroscopy confirmed that high BzMA monomer conversions (>97%) were achieved within 5 h, while GPC studies indicated that reasonably good blocking efficiencies and relatively low diblock copolymer polydispersities (Mw/Mn < 1.30) were obtained in most cases. Compared to prior literature reports, this all-methacrylic PISA formulation is particularly novel because: (i) it is the first time that higher order morphologies (e.g. worms and vesicles) have been accessed in non-polar solvents and (ii) such diblock copolymer nano-objects are expected to have potential boundary lubrication applications for engine oils.
- 60Dan, M.; Huo, F.; Zhang, X.; Wang, X.; Zhang, W. J. Polym. Sci., Part A: Polym. Chem. 2013, 51, 1573– 1584 DOI: 10.1002/pola.26527Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtFSlsg%253D%253D&md5=e7d3dcc4b664e459d97fd5b1321b1ab5Dispersion RAFT polymerization of 4-vinylpyridine in toluene mediated with the macro-RAFT agent of polystyrene dithiobenzoate: Effect of the macro-RAFT agent chain length and growth of the block copolymer nano-objectsDan, Meihan; Huo, Fei; Zhang, Xu; Wang, Xiaohui; Zhang, WangqingJournal of Polymer Science, Part A: Polymer Chemistry (2013), 51 (7), 1573-1584CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)The dispersion reversible addn.-fragmentation chain transfer (RAFT) polymn. of 4-vinylpyridine in toluene in the presence of the polystyrene dithiobenzoate (PS-CTA) macro-RAFT agent with different chain length is discussed. The RAFT polymn. undergoes an initial slow homogeneous polymn. and a subsequent fast heterogeneous one. The RAFT polymn. rate is dependent on the PS-CTA chain length, and short PS-CTA generally leads to fast RAFT polymn. The dispersion RAFT polymn. induces the self-assembly of the in situ synthesized polystyrene-b-poly(4-vinylpyridine) block copolymer into highly concd. block copolymer nano-objects. The PS-CTA chain length exerts great influence on the particle nucleation and the size and morphol. of the block copolymer nano-objects. It is found, short PS-CTA leads to fast particle nucleation and tends to produce large-sized vesicles or large-compd. micelles, and long PS-CTA leads to formation of small-sized nanospheres. Comparison between the polymn.-induced self-assembly and self-assembly of block copolymer in the block-selective solvent is made, and the great difference between the two methods is demonstrated. The present study is anticipated to be useful to reveal the chain extension and the particle growth of block copolymer during the RAFT polymn. under dispersion condition. © 2013 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013.
- 61Tan, J.; Rao, X.; Yang, J.; Zeng, Z. Macromolecules 2013, 46, 8441– 8448 DOI: 10.1021/ma401909aGoogle ScholarThere is no corresponding record for this reference.
- 62Shi, P.; Li, Q.; He, X.; Li, S.; Sun, P.; Zhang, W. Macromolecules 2014, 47, 7442– 7452 DOI: 10.1021/ma501598kGoogle ScholarThere is no corresponding record for this reference.
- 63Huang, C.-Q.; Pan, C.-Y. Polymer 2010, 51, 5115– 5121 DOI: 10.1016/j.polymer.2010.08.056Google ScholarThere is no corresponding record for this reference.
- 64Warren, N. J.; Mykhaylyk, O. O.; Mahmood, D.; Ryan, A. J.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 1023– 1033 DOI: 10.1021/ja410593nGoogle ScholarThere is no corresponding record for this reference.
- 65Zhao, W.; Gody, G.; Dong, S.; Zetterlund, P. B.; Perrier, S. Polym. Chem. 2014, 5, 6990– 7003 DOI: 10.1039/C4PY00855CGoogle Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhsVKnt7vE&md5=3612b5577d91990be47b88c39e830b8eOptimization of the RAFT polymerization conditions for the in situ formation of nano-objects via dispersion polymerization in alcoholic mediumZhao, Wei; Gody, Guillaume; Dong, Siming; Zetterlund, Per B.; Perrier, SebastienPolymer Chemistry (2014), 5 (24), 6990-7003CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Hydrophilic polymer brushes based on poly(ethylene glycol) Me ether acrylate (P(PEGA454)) or poly(ethylene glycol) Me ether methacrylate (P(PEGMA475)), both having a trithiocarbonate end group, were prepd. in water-dioxane (9 : 1) at 44 °C via RAFT polymn. and subsequently used in RAFT dispersion polymn. of styrene in isopropanol at 90 °C. RAFT reaction conditions were first optimized to prep. P(PEGA454) and P(PEGMA475) macro-RAFT agents at high monomer conversions (>90%) and very low fraction of dead chains (<1%). Both polymer brushes allowed the prepn. of well-defined amphiphilic diblock copolymers (P(PEGA454)-b-PS and P(PEGMA475)-b-PS) which self-assemble in situ into nano-objects with various morphologies. Using relatively long chain P(PEGA454) or P(PEGMA475) macro-RAFT agents (DP ~ 75) leads to the formation of near uniform spherical nanoparticles with diams. ranging from 30 to 140 nm, depending on the targeted DP of the PS block. In contrast, TEM and DLS studies demonstrated that using a shorter P(PEGA454) or P(PEGMA475) macro-RAFT agent (DP ~ 20) enables the formation of worm-like micelles, vesicles and large compd. vesicle morphologies in addn. to spheres. Cryo-TEM was used to confirm polymn. induced morphol. transition, rather than morphologies obtained via self-assembly driven by selective solvent or solvent evapn. during the prepn. of samples for characterization.
- 66Sugihara, S.; Armes, S. P.; Blanazs, A.; Lewis, A. L. Soft Matter 2011, 7, 10787– 10793 DOI: 10.1039/c1sm06593aGoogle ScholarThere is no corresponding record for this reference.
- 67He, W.-D.; Sun, X.-L.; Wan, W.-M.; Pan, C.-Y. Macromolecules 2011, 44, 3358– 3365 DOI: 10.1021/ma2000674Google ScholarThere is no corresponding record for this reference.
- 68Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Langmuir 2012, 28, 914– 922 DOI: 10.1021/la203991yGoogle ScholarThere is no corresponding record for this reference.
- 69Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Langmuir 2013, 29, 7416– 7424 DOI: 10.1021/la304279yGoogle ScholarThere is no corresponding record for this reference.
- 70Rieger, J.; Stoffelbach, F.; Bui, C.; Alaimo, D.; Jérôme, C.; Charleux, B. Macromolecules 2008, 41, 4065– 4068 DOI: 10.1021/ma800544vGoogle ScholarThere is no corresponding record for this reference.
- 71Zhang, W.; D’Agosto, F.; Boyron, O.; Rieger, J.; Charleux, B. Macromolecules 2012, 45, 4075– 4084 DOI: 10.1021/ma300596fGoogle ScholarThere is no corresponding record for this reference.
- 72Rieger, J.; Zhang, W.; Stoffelbach, F. O.; Charleux, B. Macromolecules 2010, 43, 6302– 6310 DOI: 10.1021/ma1009269Google ScholarThere is no corresponding record for this reference.
- 73Zhang, W.; D’Agosto, F.; Dugas, P.-Y.; Rieger, J.; Charleux, B. Polymer 2013, 54, 2011– 2019 DOI: 10.1016/j.polymer.2012.12.028Google ScholarThere is no corresponding record for this reference.
- 74Ratcliffe, L. P. D.; Blanazs, A.; Williams, C. N.; Brown, S. L.; Armes, S. P. Polym. Chem. 2014, 5, 3643– 3655 DOI: 10.1039/c4py00203bGoogle Scholar74https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXns1Wju7c%253D&md5=6d5f403dc2c766c3024f7e2ad246ea65RAFT polymerization of hydroxy-functional methacrylic monomers under heterogeneous conditions: effect of varying the core-forming blockRatcliffe, L. P. D.; Blanazs, A.; Williams, C. N.; Brown, S. L.; Armes, S. P.Polymer Chemistry (2014), 5 (11), 3643-3655CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Statistical copolymn. of a 1:1 molar ratio of a water-miscible monomer (2-hydroxyethyl methacrylate, HEMA) with a water-immiscible monomer (4-hydroxybutyl methacrylate, HBMA) has been conducted in water via reversible addn.-fragmentation chain transfer (RAFT) polymn. using a water-sol. poly(glycerol monomethacrylate) macromol. chain transfer agent (PGMA macro-CTA). In principle, such a hybrid formulation might be expected to be intermediate between RAFT dispersion polymn. and RAFT emulsion polymn. Under such circumstances, it is of particular interest to examine whether both monomers are actually consumed and, if so, whether their rates of reaction are comparable. Given the water-soly. of both the PGMA macro-CTA and the free radical azo initiator, it is perhaps counter-intuitive that the water-immiscible HBMA is initially consumed significantly faster than the water-miscible HEMA, as judged by 1H NMR studies of this copolymn. However, both comonomers are eventually almost fully consumed at 70 °C. A detailed phase diagram has been constructed for this RAFT formulation that enables reproducible syntheses of various pure copolymer morphologies, including spheres, worms and vesicles. It is emphasized that utilizing a 1:1 HEMA/HBMA molar ratio produces a core-forming statistical copolymer block that is isomeric with the poly(2-hydroxypropyl methacrylate) (PHPMA) core-forming block previously synthesized via RAFT aq. dispersion polymn. (see A. Blanazs et al., Macromols., 2012, 45, 5099-5107). Hence it is rather remarkable that the thermo-responsive behavior of PGMA-P(HBMA-stat-HEMA) statistical block copolymer worm gels differs qual. from that exhibited by PGMA-PHPMA diblock copolymer worm gels.
- 75Binauld, S.; Delafresnaye, L.; Charleux, B.; D’Agosto, F.; Lansalot, M. Macromolecules 2014, 47, 3461– 3472 DOI: 10.1021/ma402549xGoogle ScholarThere is no corresponding record for this reference.
- 76Zhang, X.; Boisson, F.; Colombani, O.; Chassenieux, C.; Charleux, B. Macromolecules 2014, 47, 51– 60 DOI: 10.1021/ma402125rGoogle ScholarThere is no corresponding record for this reference.
- 77Carlsson, L.; Fall, A.; Chaduc, I.; Wagberg, L.; Charleux, B.; Malmstrom, E.; D’Agosto, F.; Lansalot, M.; Carlmark, A. Polym. Chem. 2014, 5, 6076– 6086 DOI: 10.1039/C4PY00675EGoogle Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVylsrvI&md5=2210816d8f6fc6c301ea6d3aeb9d3593Modification of cellulose model surfaces by cationic polymer latexes prepared by RAFT-mediated surfactant-free emulsion polymerizationCarlsson, Linn; Fall, Andreas; Chaduc, Isabelle; Waagberg, Lars; Charleux, Bernadette; Malmstroem, Eva; D'Agosto, Franck; Lansalot, Muriel; Carlmark, AnnaPolymer Chemistry (2014), 5 (20), 6076-6086CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)This paper presents the successful surface modification of a model cellulose substrate by the prepn. and subsequent phys. adsorption of cationic polymer latexes. The first part of the work introduces novel charged polymer nanoparticles constituted of amphiphilic block copolymers based on cationic P(DMAEMA-co-MAA) as the hydrophilic segment, and PMMA as the hydrophobic segment. First, RAFT polymn. of DMAEMA in water was performed at pH 7, below its pKa. The simultaneous hydrolysis of DMAEMA led to the formation of a statistical copolymer incorporating mainly protonated DMAEMA units and some deprotonated methacrylic acid units at pH 7. The following step was the RAFT-mediated surfactant-free emulsion polymn. of MMA using P(DMAEMA-co-MAA) as a hydrophilic macromol. RAFT agent. The formed amphiphilic block copolymers self-assembled into cationic latex nanoparticles by polymn.-induced self-assembly. The nanoparticles were found to increase in size with increasing molar mass of the hydrophobic block. The cationic latexes were subsequently adsorbed to cellulose model surfaces in a quartz crystal microbalance equipment with dissipation. The adsorbed amt., in mg m-2, increased with increasing size of the nanoparticles. This approach allows for phys. surface modification of cellulose, utilizing a water suspension of particles for which both the surface chem. and the surface structure can be altered in a well-defined way.
- 78Zhang, X.; Cardozo, A. F.; Chen, S.; Zhang, W.; Julcour, C.; Lansalot, M.; Blanco, J.-F.; Gayet, F.; Delmas, H.; Charleux, B.; Manoury, E.; D’Agosto, F.; Poli, R. Chem. - Eur. J. 2014, 20, 15505– 15517 DOI: 10.1002/chem.201403819Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVyks77I&md5=51101379f694359686096b0fc5e880daCore-Shell Nanoreactors for Efficient Aqueous Biphasic CatalysisZhang, Xuewei; Cardozo, Andres F.; Chen, Si; Zhang, Wenjing; Julcour, Carine; Lansalot, Muriel; Blanco, Jean-Francois; Gayet, Florence; Delmas, Henri; Charleux, Bernadette; Manoury, Eric; D'Agosto, Franck; Poli, RinaldoChemistry - A European Journal (2014), 20 (47), 15505-15517CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)Water-borne phosphine-functionalized core-cross-linked micelles (CCM) consisting of a hydrophobic core and a hydrophilic shell were obtained as stable latexes by reversible addn.-fragmentation chain transfer (RAFT) in water in a one-pot, three-step process. Initial homogeneous aq.-phase copolymn. of methacrylic acid (MAA) and poly(ethylene oxide) Me ether methacrylate (PEOMA) is followed by copolymn. of styrene (S) and 4-diphenylphosphinostyrene (DPPS), yielding P(MAA-co-PEOMA)-b-P(S-co-DPPS) amphiphilic block copolymer micelles (M) by polymn.-induced self-assembly (PISA), and final micellar crosslinking with a mixt. of S and diethylene glycol dimethacrylate. The CCM were characterized by dynamic light scattering and NMR spectroscopy to evaluate size, dispersity, stability, and the swelling ability of various org. substrates. Coordination of [Rh(acac)(CO)2] (acac=acetylacetonate) to the core-confined phosphine groups was rapid and quant. The CCM and M latexes were then used, in combination with [Rh(acac)(CO)2], to catalyze the aq. biphasic hydroformylation of 1-octene, in which they showed high activity, recyclability, protection of the activated Rh center by the polymer scaffold, and low Rh leaching. The CCM latex gave slightly lower catalytic activity but significantly less Rh leaching than the M latex. A control expt. conducted in the presence of the sulfoxantphos ligand pointed to the action of the CCM as catalytic nanoreactors with substrate and product transport into and out of the polymer core, rather than as a surfactant in interfacial catalysis.
- 79Ihli, J.; Bots, P.; Kulak, A.; Benning, L. G.; Meldrum, F. C. Adv. Funct. Mater. 2013, 23, 1965– 1973 DOI: 10.1002/adfm.201201742Google ScholarThere is no corresponding record for this reference.
- 80Jones, E. R.; Semsarilar, M.; Blanazs, A.; Armes, S. P. Macromolecules 2012, 45, 5091– 5098 DOI: 10.1021/ma300898eGoogle ScholarThere is no corresponding record for this reference.
- 81Blanazs, A.; Madsen, J.; Battaglia, G.; Ryan, A. J.; Armes, S. P. J. Am. Chem. Soc. 2011, 133, 16581– 16587 DOI: 10.1021/ja206301aGoogle ScholarThere is no corresponding record for this reference.
- 82Odian, G. Principles of Polymerization, 4th ed.; Wiley: 2004.Google ScholarThere is no corresponding record for this reference.
- 83Warren, N. J.; Mykhaylyk, O. O.; Ryan, A. J.; Williams, M.; Doussineau, T.; Dugourd, P.; Antoine, R.; Portale, G.; Armes, S. P. J. Am. Chem. Soc. 2015, 137, 1929– 1937 DOI: 10.1021/ja511423mGoogle ScholarThere is no corresponding record for this reference.
- 84Atkins, P.; De Paula, J. Physical Chemistry; Oxford University Press: 2006.Google ScholarThere is no corresponding record for this reference.
- 85Semsarilar, M.; Jones, E. R.; Armes, S. P. Polym. Chem. 2014, 5, 195– 203 DOI: 10.1039/C3PY01042BGoogle Scholar85https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvVKjtb7L&md5=281b363840e59c0e81811b7ab500d3beComparison of pseudo-living character of RAFT polymerizations conducted under homogeneous and heterogeneous conditionsSemsarilar, Mona; Jones, Elizabeth R.; Armes, Steven P.Polymer Chemistry (2014), 5 (1), 195-203CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)RAFT dispersion polymn. of 2,2,2-trifluoroethyl methacrylate (TFEMA) is conducted in ethanol at 70 °C using either poly(2-(dimethylamino)ethyl methacrylate) or poly(methacrylic acid) as a macromol. chain transfer agent. If the diblock copolymer nanoparticles are not too large, the small refractive index difference between the PTFEMA cores and ethanol leads to minimal light scattering. This enables the pseudo-living character of RAFT formulations conducted under soln. and dispersion polymn. conditions to be compared by monitoring the loss of RAFT chain-ends via UV-visible absorption spectroscopy. Significantly fewer chain-ends are lost during RAFT dispersion polymn., suggesting that such heterogeneous formulations have greater pseudo-living character. Moreover, 19F NMR spectroscopy provides the first direct exptl. evidence that RAFT dispersion polymn. proceeds via monomer-swollen block copolymer micelles. The relatively low refractive index of PTFEMA complicates GPC anal., leading to apparent contamination of the diblock copolymer and erroneously high polydispersities. However, this artifact can be cor. by deconvolution of the GPC curves, followed by their reconstruction using appropriate refractive indexes.
- 86Blanazs, A.; Armes, S. P.; Ryan, A. J. Macromol. Rapid Commun. 2009, 30, 267– 277 DOI: 10.1002/marc.200800713Google ScholarThere is no corresponding record for this reference.
- 87Lowenstam, H. A.; Weiner, S. On Biomineralization; Oxford University Press: New York, 1989.Google ScholarThere is no corresponding record for this reference.
- 88Meldrum, F. C.; Cölfen, H. Chem. Rev. 2008, 108, 4332– 4432 DOI: 10.1021/cr8002856Google ScholarThere is no corresponding record for this reference.
- 89Berman, A.; Hanson, J.; Leiserowitz, L.; Koetzle, T. F.; Weiner, S.; Addadi, L. Science 1993, 259, 776– 779 DOI: 10.1126/science.259.5096.776Google ScholarThere is no corresponding record for this reference.
- 90Kulak, A. N.; Iddon, P.; Li, Y.; Armes, S. P.; Cölfen, H.; Paris, O.; Wilson, R. M.; Meldrum, F. C. J. Am. Chem. Soc. 2007, 129, 3729– 3736 DOI: 10.1021/ja067422eGoogle ScholarThere is no corresponding record for this reference.
- 91Kim, Y.-Y.; Ganesan, K.; Yang, P.; Kulak, A. N.; Borukhin, S.; Pechook, S.; Ribeiro, L.; Kröger, R.; Eichhorn, S. J.; Armes, S. P.; Pokroy, B.; Meldrum, F. C. Nat. Mater. 2011, 10, 890– 896 DOI: 10.1038/nmat3103Google ScholarThere is no corresponding record for this reference.
- 92Kulak, A. N.; Yang, P.; Kim, Y.-Y.; Armes, S. P.; Meldrum, F. C. Chem. Commun. 2014, 50, 67– 69 DOI: 10.1039/C3CC47904HGoogle ScholarThere is no corresponding record for this reference.
- 93Kulak, A. N.; Semsarilar, M.; Kim, Y.-Y.; Ihli, J.; Fielding, L. A.; Cespedes, O.; Armes, S. P.; Meldrum, F. C. Chem. Sci. 2014, 5, 738– 743 DOI: 10.1039/C3SC52615AGoogle Scholar93https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXitVSqtb7J&md5=71fa681a2fb3d6f8c55fa12f7f6de82bOne-pot synthesis of an inorganic heterostructure: uniform occlusion of magnetite nanoparticles within calcite single crystalsKulak, Alexander N.; Semsarilar, Mona; Kim, Yi-Yeoun; Ihli, Johannes; Fielding, Lee A.; Cespedes, Oscar; Armes, Steven P.; Meldrum, Fiona C.Chemical Science (2014), 5 (2), 738-743CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A facile one-pot method is described for the formation of novel heterostructures in which inorg. nanoparticles are homogeneously distributed throughout an inorg. single crystal matrix. Our strategy uses nanoparticles functionalised with a poly(sodium 4-styrenesulfonate)-poly(methacrylic acid) [PNaStS-PMAA] diblock copolymer as a sol. crystal growth additive. This copolymer plays a no. of essential roles. The PMAA anchor block is phys. adsorbed onto the inorg. nanoparticles, while the PNaStS block acts as an electrosteric stabilizer and ensures that the nanoparticles retain their colloidal stability in the crystal growth soln. In addn., this strong acid block promotes binding to both the nanoparticles and the host crystal, which controls nanoparticle incorporation within the host crystal lattice. We show that this approach can be used to achieve encapsulation loadings of at least 12 wt% copolymer-coated magnetite particles within calcite single crystals. Transmission electron microscopy shows that these nanoparticles are uniformly distributed throughout the calcite, and that the crystal lattice retains its continuity around the embedded magnetite particles. Characterization of these calcite/magnetite nanocomposites confirmed their magnetic properties. This new exptl. approach is expected to be quite general, such that a small family of block copolymers could be used to drive the incorporation of a wide range of pre-prepd. nanoparticles into host crystals, giving intimate mixing of phases with contrasting properties, while limiting nanoparticle aggregation and migration.
- 94Borukhin, S.; Bloch, L.; Radlauer, T.; Hill, A. H.; Fitch, A. N.; Pokroy, B. Adv. Funct. Mater. 2012, 22, 4216– 4224 DOI: 10.1002/adfm.201201079Google ScholarThere is no corresponding record for this reference.
- 95Kim, Y.-Y.; Ribeiro, L.; Maillot, F.; Ward, O.; Eichhorn, S. J.; Meldrum, F. C. Adv. Mater. 2010, 22, 2082– 2086 DOI: 10.1002/adma.200903743Google ScholarThere is no corresponding record for this reference.
- 96Tomson, M. B. J. Cryst. Growth 1983, 62, 106– 112 DOI: 10.1016/0022-0248(83)90013-1Google Scholar96https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksVehtLk%253D&md5=a1801446c6161015b96f941e19a41b0dEffect of precipitation inhibitors on calcium carbonate scale formationTomson, Mason B.Journal of Crystal Growth (1983), 62 (1), 106-12CODEN: JCRGAE; ISSN:0022-0248.A model of threshold scale inhibition is proposed based on the relative diffusion rates of the CO32- and the inhibitor ions, and a criterion for the min. concn. of inhibitor, is developed and tested. The relation ZIn(In2-) > 2(CO32-), [where ZIn is the magnitude of the charge on the inhibitor ion and () denotes molar concn.] gave good agreement with exptl. values.
- 97Ketrane, R.; Saidani, B.; Gil, O.; Leleyter, L.; Baraud, F. Desalination 2009, 249, 1397– 1404 DOI: 10.1016/j.desal.2009.06.013Google ScholarThere is no corresponding record for this reference.
- 98Ning, Y.; Fielding, L. A.; Andrews, T. S.; Growney, D. J.; Armes, S. P. Sulfate-based anionic diblock copolymer nanoparticles for efficient occlusion within zinc oxide Nanoscale 2015, 7, 6691– 6702Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 59 publications.
- Arad Lang, Iryna Polishchuk, Alexander Katsman, Boaz Pokroy. Effect of Nucleotides on the Phase and Crystal Structure of Synthetic Calcium Carbonate. Crystal Growth & Design 2023, 23
(7)
, 5117-5125. https://doi.org/10.1021/acs.cgd.3c00367
- Wangmeng Hou, Jiasheng Wu, Zheqi Li, Zhen Zhang, Yi Shi, Yongming Chen. Efficient Synthesis and PISA Behavior of Molecular Bottlebrush Block Copolymers via a Grafting-From Strategy through RAFT Dispersion Polymerization. Macromolecules 2023, 56
(3)
, 824-832. https://doi.org/10.1021/acs.macromol.2c02233
- Ouassef Nahi, Alexander Broad, Alexander N. Kulak, Helen M. Freeman, Shuheng Zhang, Thomas D. Turner, Lucien Roach, Robert Darkins, Ian J. Ford, Fiona C. Meldrum. Positively Charged Additives Facilitate Incorporation in Inorganic Single Crystals. Chemistry of Materials 2022, 34
(11)
, 4910-4923. https://doi.org/10.1021/acs.chemmater.2c00097
- Jian Sun, Chang Wang, Yu-Lin Hong, Zhi-Wei Tan, Cheng-Mei Liu. Phosphine Oxide-Containing Multifunctional Polymer via RAFT Polymerization and Its High-Density Post-Polymerization Modification in Water. ACS Applied Polymer Materials 2021, 3
(6)
, 3214-3226. https://doi.org/10.1021/acsapm.1c00403
- Yin Ning, Steven P. Armes. Efficient Occlusion of Nanoparticles within Inorganic Single Crystals. Accounts of Chemical Research 2020, 53
(6)
, 1176-1186. https://doi.org/10.1021/acs.accounts.0c00103
- Giulia Magnabosco, Iryna Polishchuk, Francesco Palomba, Enrico Rampazzo, Luca Prodi, Joanna Aizenberg, Boaz Pokroy, Giuseppe Falini. Effect of Surface Chemistry on Incorporation of Nanoparticles within Calcite Single Crystals. Crystal Growth & Design 2019, 19
(8)
, 4429-4435. https://doi.org/10.1021/acs.cgd.9b00208
- Yin Ning, Lijuan Han, Marcel Douverne, Nicholas J. W. Penfold, Matthew J. Derry, Fiona C. Meldrum, Steven P. Armes. What Dictates the Spatial Distribution of Nanoparticles within Calcite?. Journal of the American Chemical Society 2019, 141
(6)
, 2481-2489. https://doi.org/10.1021/jacs.8b12291
- Juliette Lomège, Vincent Lapinte, Claire Negrell, Jean-Jacques Robin, Sylvain Caillol. Fatty Acid-Based Radically Polymerizable Monomers: From Novel Poly(meth)acrylates to Cutting-Edge Properties. Biomacromolecules 2019, 20
(1)
, 4-26. https://doi.org/10.1021/acs.biomac.8b01156
- Jianbo Tan, Qin Xu, Yuxuan Zhang, Chundong Huang, Xueliang Li, Jun He, Li Zhang. Room Temperature Synthesis of Self-Assembled AB/B and ABC/BC Blends by Photoinitiated Polymerization-Induced Self-Assembly (Photo-PISA) in Water. Macromolecules 2018, 51
(18)
, 7396-7406. https://doi.org/10.1021/acs.macromol.8b01456
- Coit T. Hendley IV, Lee A. Fielding, Elizabeth R. Jones, Anthony J. Ryan, Steven P. Armes, Lara A. Estroff. Mechanistic Insights into Diblock Copolymer Nanoparticle–Crystal Interactions Revealed via in Situ Atomic Force Microscopy. Journal of the American Chemical Society 2018, 140
(25)
, 7936-7945. https://doi.org/10.1021/jacs.8b03828
- Yuan Zhang, Guang Han, Mengjiao Cao, Tianying Guo, Wangqing Zhang. Influence of Solvophilic Homopolymers on RAFT Polymerization-Induced Self-Assembly. Macromolecules 2018, 51
(11)
, 4397-4406. https://doi.org/10.1021/acs.macromol.8b00690
- Mehran Tehrani, Thusitha Nanda Bandara Etampawala, Mark Dadmun, and Mohammad Abir Hossain . Resolving Hierarchical Structures in Carbon Nanotube Networks Using Small- and Ultrasmall-Angle Neutron Scattering. The Journal of Physical Chemistry C 2017, 121
(40)
, 22442-22451. https://doi.org/10.1021/acs.jpcc.7b06114
- Wei Zhao, Hang T. Ta, Cheng Zhang, and Andrew K. Whittaker . Polymerization-Induced Self-Assembly (PISA) - Control over the Morphology of 19F-Containing Polymeric Nano-objects for Cell Uptake and Tracking. Biomacromolecules 2017, 18
(4)
, 1145-1156. https://doi.org/10.1021/acs.biomac.6b01788
- Jianbo Tan, Chundong Huang, Dongdong Liu, Xueliang Li, Jun He, Qin Xu, and Li Zhang . Polymerization-Induced Self-Assembly of Homopolymer and Diblock Copolymer: A Facile Approach for Preparing Polymer Nano-Objects with Higher-Order Morphologies. ACS Macro Letters 2017, 6
(3)
, 298-303. https://doi.org/10.1021/acsmacrolett.7b00134
- Juan Wang, Boxuan Li, Xing Wang, Fei Yang, Hong Shen, and Decheng Wu . Morphological Evolution of Self-Assembled Structures Induced by the Molecular Architecture of Supra-Amphiphiles. Langmuir 2016, 32
(51)
, 13706-13715. https://doi.org/10.1021/acs.langmuir.6b03550
- Pengcheng Yang, Oleksandr O. Mykhaylyk, Elizabeth R. Jones, and Steven P. Armes . RAFT Dispersion Alternating Copolymerization of Styrene with N-Phenylmaleimide: Morphology Control and Application as an Aqueous Foam Stabilizer. Macromolecules 2016, 49
(18)
, 6731-6742. https://doi.org/10.1021/acs.macromol.6b01563
- Yin Ning, Lee A. Fielding, Liam P. D. Ratcliffe, Yun-Wei Wang, Fiona C. Meldrum, and Steven P. Armes . Occlusion of Sulfate-Based Diblock Copolymer Nanoparticles within Calcite: Effect of Varying the Surface Density of Anionic Stabilizer Chains. Journal of the American Chemical Society 2016, 138
(36)
, 11734-11742. https://doi.org/10.1021/jacs.6b05563
- Sarah L. Canning, Gregory N. Smith, and Steven P. Armes . A Critical Appraisal of RAFT-Mediated Polymerization-Induced Self-Assembly. Macromolecules 2016, 49
(6)
, 1985-2001. https://doi.org/10.1021/acs.macromol.5b02602
- Yin Ning, Lee A. Fielding, Kay E. B. Doncom, Nicholas J. W. Penfold, Alexander N. Kulak, Hideki Matsuoka, and Steven P. Armes . Incorporating Diblock Copolymer Nanoparticles into Calcite Crystals: Do Anionic Carboxylate Groups Alone Ensure Efficient Occlusion?. ACS Macro Letters 2016, 5
(3)
, 311-315. https://doi.org/10.1021/acsmacrolett.6b00022
- Xuemei Liu, Peng Wang, Mengting Wan, Songjun Li, Yuan Zhang. Preparation and Application of Nitrobenzene-Containing Copolymer Nanoparticles Obtained by RAFT Dispersion Polymerization. Journal of Inorganic and Organometallic Polymers and Materials 2024, 34
(8)
, 3647-3659. https://doi.org/10.1007/s10904-024-03041-3
- Bing Cao, Jian Zhang, Yingfei Ma, Yanan Wang, Yun li, Ruixue Wang, Donghai Cao, Yongzhen Yang, Ruiping Zhang. Dual‐Polymer Functionalized Melanin‐AgNPs Nanocomposite with Hydroxyapatite Binding Ability to Penetrate and Retain in Biofilm Sequentially Treating Periodontitis. Small 2024, https://doi.org/10.1002/smll.202400771
- Bingfeng Shi, Wanchao Hu, Shiyi Li, Zhinan Xia, Changli Lü. Perovskite photoinitiated RAFT-mediated polymerization-induced self-assembly for organic–inorganic hybrid nanomaterials. Inorganic Chemistry Frontiers 2024, 11
(8)
, 2471-2478. https://doi.org/10.1039/D4QI00233D
- Jiahao Zhang, Biao Xiong, Ziyu Fu, Yin Ning, Dan Li. Synergistic Effect of Hydroxyl and Carboxyl Groups on Promoting Nanoparticle Occlusion within Calcite. Small 2023, 19
(18)
https://doi.org/10.1002/smll.202207843
- Ruiling Du, Lee A. Fielding. Preparation of polymer nanoparticle-based complex coacervate hydrogels using polymerisation-induced self-assembly derived nanogels. Soft Matter 2023, 19
(11)
, 2074-2081. https://doi.org/10.1039/D2SM01534J
- Jian Sun, Yu-Lin Hong, Xiu-Qin Fang, Chang Wang, Cheng-Mei Liu. Fluorescent phosphine oxide-containing hyperbranched polyesters: design, synthesis and their application for Fe
3+
detection. Journal of Materials Chemistry C 2023, 11
(5)
, 1927-1936. https://doi.org/10.1039/D2TC04608C
- Bartosz Marzec, Jessica Walker, Yasmeen Jhons, Fiona C. Meldrum, Michael Shaver, Fabio Nudelman. Micron-sized biogenic and synthetic hollow mineral spheres occlude additives within single crystals. Faraday Discussions 2022, 235 , 536-550. https://doi.org/10.1039/D1FD00095K
- Sètuhn Jimaja, Spyridon Varlas, Jeffrey C. Foster, Daniel Taton, Andrew P. Dove, Rachel K. O'Reilly. Stimuli-responsive and core cross-linked micelles developed by NiCCo-PISA of helical poly(aryl isocyanide)s. Polymer Chemistry 2022, 13
(27)
, 4047-4053. https://doi.org/10.1039/D2PY00397J
- Yin Ning, Steven P. Armes, Dan Li. Polymer‐Inorganic Crystalline Nanocomposite Materials via Nanoparticle Occlusion. Macromolecular Rapid Communications 2022, 43
(14)
https://doi.org/10.1002/marc.202100793
- Jie Ren, Yujing Liu, Hanying Li. Incorporating polymers within a single‐crystal: From heterogeneous structure to multiple functions. Journal of Polymer Science 2022, 60
(7)
, 1151-1173. https://doi.org/10.1002/pol.20210408
- Fanny Coumes, François Stoffelbach, Jutta Rieger. Polymerization‐Induced Self‐Assembly: From Macromolecular Engineering Toward Applications. 2022, 1-84. https://doi.org/10.1002/9783527815562.mme0037
- Hui Wang, Christophe Fliedel, Eric Manoury, Rinaldo Poli. Core-crosslinked micelles with a poly-anionic poly(styrene sulfonate)-based outer shell made by RAFT polymerization. Polymer 2022, 243 , 124640. https://doi.org/10.1016/j.polymer.2022.124640
- Matthew Boon, Franca Jones. Phosphonate inhibitors: Types, solution chemistry, and applications. 2022, 697-722. https://doi.org/10.1016/B978-0-12-822896-8.00018-2
- Wen-Li Wang, Ren-Hua Jin. Synthesis and self-assembly of amphiphilic comb-copolymers possessing polyethyleneimine and its derivatives: Site-selective formation of loop-cluster covered vesicles and flower micelles. Polymer 2021, 212 , 123289. https://doi.org/10.1016/j.polymer.2020.123289
- Yasemin Samav, Bernice Akpinar, Gökhan Kocak, Vural Bütün. Preparation of Responsive Zwitterionic Diblock Copolymers Containing Phosphate and Phosphonate Groups. Macromolecular Research 2020, 28
(12)
, 1134-1141. https://doi.org/10.1007/s13233-020-8148-3
- Yin Ning, Yide Han, Lijuan Han, Matthew J. Derry, Steven P. Armes. Exerting Spatial Control During Nanoparticle Occlusion within Calcite Crystals. Angewandte Chemie International Edition 2020, 59
(41)
, 17966-17973. https://doi.org/10.1002/anie.202007110
- Yin Ning, Yide Han, Lijuan Han, Matthew J. Derry, Steven P. Armes. Exerting Spatial Control During Nanoparticle Occlusion within Calcite Crystals. Angewandte Chemie 2020, 132
(41)
, 18122-18129. https://doi.org/10.1002/ange.202007110
- Arad Lang, Iryna Polishchuk, Eva Seknazi, Jochen Feldmann, Alexander Katsman, Boaz Pokroy. Bioinspired Molecular Bridging in a Hybrid Perovskite Leads to Enhanced Stability and Tunable Properties. Advanced Functional Materials 2020, 30
(42)
, 2005136. https://doi.org/10.1002/adfm.202005136
- Taeyoon Kim, Joo-Hyun Song, Jong-Ho Back, Bongkuk Seo, Choong-Sun Lim, Hyun-Jong Paik, Wonjoo Lee. Flame Retardant Submicron Particles via Surfactant-Free RAFT Emulsion Polymerization of Styrene Derivatives Containing Phosphorous. Polymers 2020, 12
(6)
, 1244. https://doi.org/10.3390/polym12061244
- Giulia Magnabosco, Hagar Hauzer, Simona Fermani, Matteo Calvaresi, Franco Corticelli, Meganne Christian, Cristiano Albonetti, Vittorio Morandi, Jonathan Erez, Giuseppe Falini. Bionic synthesis of a magnetic calcite skeletal structure through living foraminifera. Materials Horizons 2019, 6
(9)
, 1862-1867. https://doi.org/10.1039/C9MH00495E
- Lee A. Fielding, Coit T. Hendley IV, Emily Asenath-Smith, Lara A. Estroff, Steven P. Armes. Rationally designed anionic diblock copolymer worm gels are useful model systems for calcite occlusion studies. Polymer Chemistry 2019, 10
(37)
, 5131-5141. https://doi.org/10.1039/C9PY00889F
- Xiaolong He, Xuefeng Li, Jinfeng Dong. Self-assembly of well-defined amphiphilic poly(N-(2-methacryloylxyethyl)pyrrolidone)- poly(lauryl methacrylate) diblock copolymers in non-polar solvent. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2019, 577 , 493-499. https://doi.org/10.1016/j.colsurfa.2019.05.095
- Dan Li, Meng Huo, Lei Liu, Min Zeng, Xi Chen, Xiaosong Wang, Jinying Yuan. Overcoming Kinetic Trapping for Morphology Evolution during Polymerization‐Induced Self‐Assembly. Macromolecular Rapid Communications 2019, 40
(16)
https://doi.org/10.1002/marc.201900202
- Marcel Douverne, Yin Ning, Aikaterini Tatani, Fiona C. Meldrum, Steven P. Armes. How Many Phosphoric Acid Units Are Required to Ensure Uniform Occlusion of Sterically Stabilized Nanoparticles within Calcite?. Angewandte Chemie International Edition 2019, 58
(26)
, 8692-8697. https://doi.org/10.1002/anie.201901307
- Marcel Douverne, Yin Ning, Aikaterini Tatani, Fiona C. Meldrum, Steven P. Armes. How Many Phosphoric Acid Units Are Required to Ensure Uniform Occlusion of Sterically Stabilized Nanoparticles within Calcite?. Angewandte Chemie 2019, 131
(26)
, 8784-8789. https://doi.org/10.1002/ange.201901307
- Yin Ning, Lee A. Fielding, John Nutter, Alexander N. Kulak, Fiona C. Meldrum, Steven P. Armes. Spatially Controlled Occlusion of Polymer‐Stabilized Gold Nanoparticles within ZnO. Angewandte Chemie 2019, 131
(13)
, 4346-4351. https://doi.org/10.1002/ange.201814492
- Yin Ning, Lee A. Fielding, John Nutter, Alexander N. Kulak, Fiona C. Meldrum, Steven P. Armes. Spatially Controlled Occlusion of Polymer‐Stabilized Gold Nanoparticles within ZnO. Angewandte Chemie International Edition 2019, 58
(13)
, 4302-4307. https://doi.org/10.1002/anie.201814492
- Qinghua Xu, Chun Tian, Lifen Zhang, Zhenping Cheng, Xiulin Zhu. Photo‐Controlled Polymerization‐Induced Self‐Assembly (Photo‐PISA): A Novel Strategy Using In Situ Bromine‐Iodine Transformation Living Radical Polymerization. Macromolecular Rapid Communications 2019, 40
(2)
https://doi.org/10.1002/marc.201800327
- Jun He, Qin Xu, Jianbo Tan, Li Zhang. Ketone‐Functionalized Polymer Nano‐Objects Prepared via Photoinitiated Polymerization‐Induced Self‐Assembly (Photo‐PISA) Using a Poly(diacetone acrylamide)‐Based Macro‐RAFT Agent. Macromolecular Rapid Communications 2019, 40
(2)
https://doi.org/10.1002/marc.201800296
- Dao Le, Dominic Keller, Guillaume Delaittre. Reactive and Functional Nanoobjects by Polymerization‐Induced Self‐Assembly. Macromolecular Rapid Communications 2019, 40
(2)
https://doi.org/10.1002/marc.201800551
- Pallabi Samaddar, Akash Deep, Ki-Hyun Kim. An engineering insight into block copolymer self-assembly: Contemporary application from biomedical research to nanotechnology. Chemical Engineering Journal 2018, 342 , 71-89. https://doi.org/10.1016/j.cej.2018.01.062
- Jianbo Tan, Qin Xu, Xueliang Li, Jun He, Yuxuan Zhang, Xiaocong Dai, Liangliang Yu, Ruiming Zeng, Li Zhang. Enzyme‐PISA: An Efficient Method for Preparing Well‐Defined Polymer Nano‐Objects under Mild Conditions. Macromolecular Rapid Communications 2018, 39
(9)
https://doi.org/10.1002/marc.201700871
- Jie Ma, Henitsoa M. Andriambololona, Damien Quemener, Mona Semsarilar. Membrane preparation by sequential spray deposition of polymer PISA nanoparticles. Journal of Membrane Science 2018, 548 , 42-49. https://doi.org/10.1016/j.memsci.2017.11.010
- Kazuho Nakamura, Taichi Takagishi. Change in Zeta Potential of Calcium Carbonate during Reactive Crystallization. KAGAKU KOGAKU RONBUNSHU 2018, 44
(1)
, 29-34. https://doi.org/10.1252/kakoronbunshu.44.29
- Jianbo Tan, Dongdong Liu, Chundong Huang, Xueliang Li, Jun He, Qin Xu, Li Zhang. Photoinitiated Polymerization‐Induced Self‐Assembly of Glycidyl Methacrylate for the Synthesis of Epoxy‐Functionalized Block Copolymer Nano‐Objects. Macromolecular Rapid Communications 2017, 38
(15)
https://doi.org/10.1002/marc.201700195
- Jonathan Yeow, Cyrille Boyer. Photoinitiated Polymerization-Induced Self-Assembly (Photo-PISA): New Insights and Opportunities. Advanced Science 2017, 4
(7)
, 1700137. https://doi.org/10.1002/advs.201700137
- Lewis D. Blackman, Kay E. B. Doncom, Matthew I. Gibson, Rachel K. O'Reilly. Comparison of photo- and thermally initiated polymerization-induced self-assembly: a lack of end group fidelity drives the formation of higher order morphologies. Polymer Chemistry 2017, 8
(18)
, 2860-2871. https://doi.org/10.1039/C7PY00407A
- Andrew B. Lowe. RAFT alcoholic dispersion polymerization with polymerization-induced self-assembly. Polymer 2016, 106 , 161-181. https://doi.org/10.1016/j.polymer.2016.08.082
- Ilya Captain, Timothy J. Deming. Methionine sulfoxide and phosphonate containing double hydrophilic block copolypeptides and their mineralization of calcium carbonate. Journal of Polymer Science Part A: Polymer Chemistry 2016, 54
(23)
, 3707-3712. https://doi.org/10.1002/pola.28264
- Yuming Zhao, Wen Zhu, Ying Wu, Lin Qu, Zhengping Liu, Ke Zhang. An aggregation-induced emission star polymer with pH and metal ion responsive fluorescence. Polymer Chemistry 2016, 7
(42)
, 6513-6520. https://doi.org/10.1039/C6PY01488G
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 98 other publications.
- 1Lupa, L.; Popa, A.; Ilia, G. In Phosphorus-Based Polymers: From Synthesis to Applications; The Royal Society of Chemistry: 2014; pp 225– 251.There is no corresponding record for this reference.
- 2Clearfield, A. Curr. Opin. Solid State Mater. Sci. 2002, 6, 495– 506 DOI: 10.1016/S1359-0286(02)00151-12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXktVKgtLc%253D&md5=6d83bab221755d65704b8d80dfbdb213Recent advances in metal phosphonate chemistry. IIClearfield, AbrahamCurrent Opinion in Solid State & Materials Science (2003), 6 (6), 495-506CODEN: COSSFX; ISSN:1359-0286. (Elsevier Science Ltd.)A review. Phosphonic acids are more acidic than carboxylic acids. They are nonplanar and possess twice the no. of protons. These attributes lead to strong H bond formation that can be used in design of crystal-engineered structures and self-assembled supramol. arrays. Their superior cation complexing ability gives a vast no. of metal complexes and their use in lanthanide and actinide sepns. Recent work, largely drawn from the work of the author's research group, is summarized in this update of the author's 1996 review [Curr Opin Solid State Mater Sci 1(1996) 268].
- 3van den Brand, J.; Van Gils, S.; Beentjes, P. C. J.; Terryn, H.; Sivel, V.; de Wit, J. H. W. Prog. Org. Coat. 2004, 51, 339– 350 DOI: 10.1016/j.porgcoat.2004.08.0053https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVaitLfO&md5=d3624facd5be90053eb453a201fddac0Improving the adhesion between epoxy coatings and aluminum substratesvan den Brand, J.; Van Gils, S.; Beentjes, P. C. J.; Terryn, H.; Sivel, V.; de Wit, J. H. W.Progress in Organic Coatings (2004), 51 (4), 339-350CODEN: POGCAT; ISSN:0300-9440. (Elsevier B.V.)Two methods were developed to improve adhesion over the life of a commonly used epoxy coating on an aluminum substrate. The first method involves application of a 10 nm polymeric layer on the aluminum substrate prior to application of the epoxy coating. The functional groups of the polymer undergo chemisorption onto the oxide surface and are involved in curing of the epoxy resin. The polymers are poly(acrylic acid) (PAA), poly(ethylene-alt-maleic anhydride) (PEMah) and poly(vinyl phosphonic acid) (PvPA). The interface of the epoxy coating and aluminum showed that the polymeric layer was indeed part of the cured product. For poly(ethylene-alt-maleic anhydride), there was a clear region between coating and substrate while for the two other polymers, a weakly cured interface region was formed. Adhesion and durability were also improved by hydration of the aluminum substrate by immersion in boiling water, resulting in formation of a porous pseudo-boehmite oxyhydroxide layer. The epoxy coating fully penetrated into this layer. The adhesion of the epoxy coatings was tested initially and after exposure to 40° water and 40° 5% aq. acetic acid. The poly(ethylene-alt-maleic anhydride)-based system showed good initial adhesion and durability on exposure to water, while the systems based on the other two polymers did not. The pseudoboehmite-based system also showed good initial adhesion and durability in the presence of water. None of the proposed systems were able to withstand 40° 5% acetic acid and severe corrosion of Al occurred underneath the epoxy coating.
- 4Zhang, Q. M.; Serpe, M. J. Macromolecules 2014, 47, 8018– 8025 DOI: 10.1021/ma501997xThere is no corresponding record for this reference.
- 5Alexandrino, E. M.; Ritz, S.; Marsico, F.; Baier, G.; Mailander, V.; Landfester, K.; Wurm, F. R. J. Mater. Chem. B 2014, 2, 1298– 1306 DOI: 10.1039/c3tb21295e5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXisFSitro%253D&md5=1b741a14fff5f826b14b79d89053ca76Paclitaxel-loaded polyphosphate nanoparticles: a potential strategy for bone cancer treatmentAlexandrino, Evandro M.; Ritz, Sandra; Marsico, Filippo; Baier, Grit; Mailaender, Volker; Landfester, Katharina; Wurm, Frederik R.Journal of Materials Chemistry B: Materials for Biology and Medicine (2014), 2 (10), 1298-1306CODEN: JMCBDV; ISSN:2050-7518. (Royal Society of Chemistry)While it has been shown that phosphates can target mols. and nanocarriers to bone we herein demonstrate the prepn. of polyphosphate nanoparticles loaded with paclitaxel using a simple miniemulsion/solvent-evapn. technique as a model for chemotherapeutic delivery. Polyphosphates exhibit much higher structural versatility, relying on the pentavalence of the phosphorus center compared to conventional polyesters. This versatility allows for the development of new degradable polymeric carriers with inherent bone adhesion ability by the interaction of the nanoparticles with a calcium phosphate material used for bone regeneration. The novel polyphosphate nanoparticles were investigated in detail with respect to their size distribution, zeta-potential, thermal and morphol. properties and were further proven to be efficiently loaded with a hydrophobic drug (up to 15 wt%). The in vitro cytotoxicity was assessed against human cancer cell lines (HeLa and Saos-2), and the paclitaxel-loaded nanoparticles showed a similar cytotoxicity profile similar to the com. available formulation Taxomedac and the pure paclitaxel for loading ratios of 10 wt% but addnl. proved efficient adhesion on calcium phosphate granules allowing drug delivery to bone. This first report demonstrates that polyphosphate nanoparticles are promising materials for the development of systemic or local bone cancer treatment, even by direct application or by formation of composites with calcium phosphate cements.
- 6Brunklaus, G.; Schauff, S.; Markova, D.; Klapper, M.; Müllen, K.; Spiess, H.-W. J. Phys. Chem. B 2009, 113, 6674– 6681 DOI: 10.1021/jp901714fThere is no corresponding record for this reference.
- 7Quittmann, U.; Lecamp, L.; El Khatib, W.; Youssef, B.; Bunel, C. Macromol. Chem. Phys. 2001, 202, 628– 635 DOI: 10.1002/1521-3935(20010301)202:5<628::AID-MACP628>3.0.CO;2-TThere is no corresponding record for this reference.
- 8Canadell, J.; Hunt, B. J.; Cook, A. G.; Mantecón, A.; Cádiz, V. Polym. Degrad. Stab. 2007, 92, 1482– 1490 DOI: 10.1016/j.polymdegradstab.2007.05.0178https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXos1Sis70%253D&md5=42925672d0dbcdd3261acdd9d2d44200Flame retardance and shrinkage reduction of polystyrene modified with acrylate-containing phosphorus and crosslinkable spiro-orthoester moietiesCanadell, J.; Hunt, B. J.; Cook, A. G.; Mantecon, A.; Cadiz, V.Polymer Degradation and Stability (2007), 92 (8), 1482-1490CODEN: PDSTDW; ISSN:0141-3910. (Elsevier Ltd.)Styrene was radically copolymd. with a spiro-orthoester with an acrylate group (SOE-AC), and terpolymd. with SOE-AC and diethyl(methacryloyloxymethyl)phosphonate (DEMMP). This was done for several different feed ratios, to obtain polymers with spiro-orthoester moieties in the side chain. These polymers were then crosslinked with ytterbium triflate, as cationic initiator, via the double ring-opening polymn. The thermal stability and fire retardant properties of these materials were evaluated by TGA and LOI. The DEMMP-contg. polymers give materials which were significantly more flame retardant than the nonphosphorus-contg. materials, as indicated by the LOI measurements. The vol. changes measured upon crosslinking of the polymers were evaluated by d. measurements with a gas pycnometer. In all the cases, expansion was obsd. This indicates that SOE-AC is an effective monomer for crosslinkable polymers without vol. changes.
- 9Köppl, T.; Brehme, S.; Pospiech, D.; Fischer, O.; Wolff-Fabris, F.; Altstädt, V.; Schartel, B.; Döring, M. J. Appl. Polym. Sci. 2013, 128, 3315– 3324 DOI: 10.1002/app.38520There is no corresponding record for this reference.
- 10Monge, S.; Canniccioni, B.; Graillot, A.; Robin, J.-J. Biomacromolecules 2011, 12, 1973– 1982 DOI: 10.1021/bm2004803There is no corresponding record for this reference.
- 11Lin, Z.; Cao, S.; Chen, X.; Wu, W.; Li, J. Biomacromolecules 2013, 14, 2206– 2214 DOI: 10.1021/bm4003442There is no corresponding record for this reference.
- 12David, G.; Asri, Z. E.; Rich, S.; Castignolles, P.; Guillaneuf, Y.; Lacroix-Desmazes, P.; Boutevin, B. Macromol. Chem. Phys. 2009, 210, 631– 639 DOI: 10.1002/macp.200800540There is no corresponding record for this reference.
- 13Canniccioni, B.; Monge, S.; David, G.; Robin, J.-J. Polym. Chem. 2013, 4, 3676– 3685 DOI: 10.1039/c3py00426k13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXovFWktb0%253D&md5=ca49c191735fec9f1aac492dc62b29b2RAFT polymerization of dimethyl(methacryloyloxy)methyl phosphonate and its phosphonic acid derivative: a new opportunity for phosphorus-based materialsCanniccioni, Benjamin; Monge, Sophie; David, Ghislain; Robin, Jean-JacquesPolymer Chemistry (2013), 4 (13), 3676-3685CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)The RAFT polymn. of dimethyl(methacryloyloxy)methyl phosphonate (MAPC1) using dithioester chain transfer agents is reported for the first time to our knowledge in the literature. Poly(dimethyl(methacryloyloxy)methyl phosphonate) (PMAPC1) was synthesized in DMF at 70° with a good control over the mol. wt., the latter ranging from 8000 to 24 000 g mol-1. Polymers were characterized by 1H and 31P NMR, and size exclusion chromatog. using both refractive index and triple detection. The effect of the solvent was also investigated. We demonstrated that polarity had an effect on the control of the polymn. as low polarity led to low polymn. rate and termination reactions whereas very high polarity resulted in high polymn. rate but also transfer reactions. Addnl., we showed that it was possible to restart the RAFT polymn. from a PMAPC1 macrochain transfer agent. PMAPC1 was hydrolyzed to afford poly((methacryloyloxy)methyl phosphonic acid) (hPMAPC1), whose pKa values were detd. Finally, we achieved the controlled RAFT polymn. of the (methacryloyloxy)methyl phosphonic acid (hMAPC1). The easy obtaining of PMAPC1 and hPMAPC1 opens the way to the synthesis of complex polymer architectures.
- 14Graillot, A.; Monge, S.; Faur, C.; Bouyer, D.; Robin, J.-J. Polym. Chem. 2013, 4, 795– 803 DOI: 10.1039/C2PY20720F14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtl2nsQ%253D%253D&md5=4bd338a8e77928e0045f58ec2816a9edSynthesis by RAFT of innovative well-defined (co)polymers from a novel phosphorus-based acrylamide monomerGraillot, Alain; Monge, Sophie; Faur, Catherine; Bouyer, Denis; Robin, Jean-JacquesPolymer Chemistry (2013), 4 (3), 795-803CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)The present contribution reports on the synthesis and controlled polymn. of a novel acrylamide monomer contg. phosphonated moieties, namely diethyl-2-(acrylamido)ethylphosphonate. This monomer appears to be of great interest due to the phosphonated moieties, which can lead to a wide range of applications, assocd. with the chem. stability of the acrylamide compared to more common (meth)acrylate monomers. Reversible Addn.-Fragmentation Transfer (RAFT) polymn. of this monomer was studied using two different trithiocarbonate chain transfer agents, and it allowed the synthesis of poly(diethyl-2-(acrylamido)ethylphosphonate) with controlled mol. wt. and low dispersity. Addnl., a diblock copolymer was successfully prepd. by a similar RAFT procedure using thermosensitive poly(N-n-propylacrylamide) as a macro-chain transfer agent. A combination of both stimuli-responsive and phosphonated ester or phosphonic diacid (after hydrolysis) contg. blocks appears valuable for drug delivery or water treatment, for instance.
- 15Grøndahl, L.; Suzuki, S.; Wentrup-Byrne, E. Chem. Commun. 2008, 3314– 3316 DOI: 10.1039/b803223hThere is no corresponding record for this reference.
- 16Miyatake, K.; Hay, A. S. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 1854– 1859 DOI: 10.1002/pola.116316https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXjsVaitLo%253D&md5=d4141b680795dc56cc6be9b3fe953aecSynthesis of novel phosphinic acid-containing polymersMiyatake, Kenji; Hay, Allan S.Journal of Polymer Science, Part A: Polymer Chemistry (2001), 39 (11), 1854-1859CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)Three arylene difluoride monomers contg. phosphine oxide (1), phosphinic acid (2), or phosphinate ester (3) groups were prepd. and polymd. with bisphenol A to give novel poly(arylene ether)s (4, 5, and 6). The polymers obtained had moderate mol. wts. (ηinh: 0.14-0.30 dL g-1 in N-methylpyrrolidinone) and glass-transition temps. (Tg: 102-200°), depending on the phosphine group in the main chain. Using bis(4-fluorophenyl)sulfone as a comonomer improved the polymn. to give copolymers with higher soln. viscosities. The stoichiometric investigation revealed that 7 mol % excess of fluoride monomer gave the highest mol. wt. copolymer 8 with ηinh of 0.78 dL g-1, which had a Tg of 176 °C, a Td10% of 432°, and formed a hard film by casting from soln.
- 17Watson, J. D.; Crick, F. H. C. Nature 1974, 248, 765– 765 DOI: 10.1038/248765a0There is no corresponding record for this reference.
- 18Ornelas-Megiatto, C.; Wich, P. R.; Fréchet, J. M. J. J. Am. Chem. Soc. 2012, 134, 1902– 1905 DOI: 10.1021/ja207366kThere is no corresponding record for this reference.
- 19Uchiyama, T.; Kiritoshi, Y.; Watanabe, J.; Ishihara, K. Biomaterials 2003, 24, 5183– 5190 DOI: 10.1016/S0142-9612(03)00441-1There is no corresponding record for this reference.
- 20Pegoraro, C.; Cecchin, D.; Gracia, L. S.; Warren, N.; Madsen, J.; Armes, S. P.; Lewis, A.; MacNeil, S.; Battaglia, G. Cancer Lett. 2013, 334, 328– 337 DOI: 10.1016/j.canlet.2013.02.00720https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXktlKksLk%253D&md5=b83b822803e34fa092adba772af91f37Enhanced drug delivery to melanoma cells using PMPC-PDPA polymersomesPegoraro, Carla; Cecchin, Denis; Gracia, Lorena Simon; Warren, Nicholas; Madsen, Jeppe; Armes, Steven P.; Lewis, Andrew; MacNeil, Sheila; Battaglia, GiuseppeCancer Letters (New York, NY, United States) (2013), 334 (2), 328-337CODEN: CALEDQ; ISSN:0304-3835. (Elsevier)We present the efficient and stable encapsulation of doxorubicin within pH sensitive polymeric vesicles (polymersomes) for intracellular and nuclear delivery to melanoma cells. We demonstrate that PMPC25-PDPA70 polymersomes can encapsulate doxorubicin for long periods of time without significant drug release. We demonstrate that empty polymersomes are non-toxic and that they are quickly and more efficiently internalised by melanoma cells compared to healthy cells. Encapsulated doxorubicin has a strong cytotoxic effect on both healthy and cancerous cells, but when encapsulated it had a preferential effect on melanoma cells indicating that this formulation can be used to achieve an enhanced drug delivery to cancerous cells rather than to the healthy surrounding cells.
- 21Hemp, S. T.; Smith, A. E.; Bryson, J. M.; Allen, M. H.; Long, T. E. Biomacromolecules 2012, 13, 2439– 2445 DOI: 10.1021/bm300689fThere is no corresponding record for this reference.
- 22Liu, G.; Jin, Q.; Liu, X.; Lv, L.; Chen, C.; Ji, J. Soft Matter 2011, 7, 662– 669 DOI: 10.1039/C0SM00708KThere is no corresponding record for this reference.
- 23Matyjaszewski, K. Macromolecules 2012, 45, 4015– 4039 DOI: 10.1021/ma300171923https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XlsVaqs7w%253D&md5=350b580bd1bb46c21ba5dbfd65a6811dAtom Transfer Radical Polymerization (ATRP): Current Status and Future PerspectivesMatyjaszewski, KrzysztofMacromolecules (Washington, DC, United States) (2012), 45 (10), 4015-4039CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A review. Current status and future perspectives in atom transfer radical polymn. (ATRP) are presented. Special emphasis is placed on mechanistic understanding of ATRP, recent synthetic and process development, and new controlled polymer architectures enabled by ATRP. New hybrid materials based on org./inorg. systems and natural/synthetic polymers are presented. Some current and forthcoming applications are described.
- 24Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2012, 65, 985– 1076 DOI: 10.1071/CH12295There is no corresponding record for this reference.
- 25Lobb, E. J.; Ma, I.; Billingham, N. C.; Armes, S. P.; Lewis, A. L. J. Am. Chem. Soc. 2001, 123, 7913– 7914 DOI: 10.1021/ja003906dThere is no corresponding record for this reference.
- 26Madsen, J.; Canton, I.; Warren, N. J.; Themistou, E.; Blanazs, A.; Ustbas, B.; Tian, X.; Pearson, R.; Battaglia, G.; Lewis, A. L.; Armes, S. P. J. Am. Chem. Soc. 2013, 135, 14863– 14870 DOI: 10.1021/ja407380tThere is no corresponding record for this reference.
- 27Yusa, S.-i.; Fukuda, K.; Yamamoto, T.; Ishihara, K.; Morishima, Y. Biomacromolecules 2005, 6, 663– 670 DOI: 10.1021/bm0495553There is no corresponding record for this reference.
- 28Ohno, S.; Hasegawa, S.; Liu, H.; Ishihara, K.; Yusa, S.-i. Polym. J. 2015, 47, 71 DOI: 10.1038/pj.2014.9228https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVGntQ%253D%253D&md5=4b23620ea73ecacb0ab49ffed37613e5Aggregation behavior in water of amphiphilic diblock copolymers bearing biocompatible phosphorylcholine and cholesteryl groupsOhno, Sayaka; Hasegawa, Shoto; Liu, Huihua; Ishihara, Kazuhiko; Yusa, Shin-ichiPolymer Journal (Tokyo, Japan) (2015), 47 (1), 71-76CODEN: POLJB8; ISSN:0032-3896. (NPG Nature Asia-Pacific)Poly(2-(methacryloyloxy)ethyl phosphorylcholine)-block-poly(cholesteryl 6-methacryloyloxyhexanoate) (PMPC82-b-PChMn) copolymers with different PChM block lengths were prepd. via reversible addn.-fragmentation chain transfer controlled/living radical polymn. using a PMPC-based macro-chain transfer agent. The subscript no. and n (=3 and 6) refer to the d.p. of the PMPC and PChM blocks, resp. PMPC82-b-PChMn cannot dissolve in water directly due to the strong hydrophobic nature of the PChM block. To prep. the aq. soln., the diblock copolymer was dissolved in an org. solvent and then dialyzed against pure water. These diblock copolymers formed spherical and rod-like micelles in water, depending on the compn. of cholesteryl (Chol) group in the polymer. The prepd. aggregates were characterized using static light scattering, dynamic light scattering, transmission electron microscopy and fluorescence probe techniques. The characterization results suggest that the morphol. of the polymer aggregates can be controlled from spherical to rod-like micelles by increasing the no. of Chol groups in the polymer.
- 29Huang, J.; Matyjaszewski, K. Macromolecules 2005, 38, 3577– 3583 DOI: 10.1021/ma047564yThere is no corresponding record for this reference.
- 30Suzuki, S.; Whittaker, M. R.; Grøndahl, L.; Monteiro, M. J.; Wentrup-Byrne, E. Biomacromolecules 2006, 7, 3178– 3187 DOI: 10.1021/bm060583qThere is no corresponding record for this reference.
- 31Blidi, I.; Geagea, R.; Coutelier, O.; Mazieres, S.; Violleau, F.; Destarac, M. Polym. Chem. 2012, 3, 609– 612 DOI: 10.1039/c2py00541g31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvFalurc%253D&md5=cbc2b11df46ed53d703cf1ef26916078Aqueous RAFT/MADIX polymerization of vinylphosphonic acidBlidi, Issam; Geagea, Roland; Coutelier, Olivier; Mazieres, Stephane; Violleau, Frederic; Destarac, MathiasPolymer Chemistry (2012), 3 (3), 609-612CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)RAFT/MADIX polymn. of vinylphosphonic acid (VPA) was controlled in water with an O-Et xanthate transfer agent. This represents the first example of reversible deactivation radical polymn. of a monomer bearing an unprotected phosphonic acid function. Hence, macromol. engineering of polyphosphonates can now be envisioned by directly polymg. VPA in water.
- 32Markova, D.; Kumar, A.; Klapper, M.; Müllen, K. Polymer 2009, 50, 3411– 3421 DOI: 10.1016/j.polymer.2009.06.011There is no corresponding record for this reference.
- 33David, G.; Negrell, C.; Manseri, A.; Boutevin, B. J. Appl. Polym. Sci. 2009, 114, 2213– 2220 DOI: 10.1002/app.3043833https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVCgtb3K&md5=a82c0421fd3862209b38be8f247f8c41Poly(MMA)-b-poly(monophosphonic acrylate) diblock copolymers obtained by ATRP and used as additives for anticorrosive coatingsDavid, Ghislain; Negrell, Claire; Manseri, Abdelatif; Boutevin, BernardJournal of Applied Polymer Science (2009), 114 (4), 2213-2220CODEN: JAPNAB; ISSN:0021-8995. (John Wiley & Sons, Inc.)Atom transfer radical polymn. (ATRP) of dimethyl(methacryloyloxy)methyl phosphonate (MAC1P) was investigated in toluene, in the presence of Me 2-bromoisobutyrate as the initiator, and using different metal and ligand systems. Polymn. proceeded with very low monomer conversion, which was attributed to the ability of phosphorus to complex the copper ions, removing copper ions from original ligand, and then stopping the MAC1P polymn. Poly(MMA)-b-poly(phosphonate acrylate) diblock copolymer structure was efficiently obtained by the ATRP process, based on a four-step reaction. Poly(MMA)-b-poly(tert-Bu acrylate) diblock copolymer was first obtained by ATRP, then the tert-Bu groups were removed and phosphonate functions were incorporated by esterification reaction, using 4-dimethylaminopyridine as the catalyst. This new diblock copolymer was used as an additive for anticorrosive coating; however, no improvement (using the salt spray test technique) was obsd. comparatively with the statistical copolymer with the same acid content. P-contg. additive led to the increased corrosion resistance. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009.
- 34Mukumoto, K.; Zhong, M.; Matyjaszewski, K. Eur. Polym. J. 2014, 56, 11– 16 DOI: 10.1016/j.eurpolymj.2014.03.02934https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpslCrsrc%253D&md5=38541cf2544ad1e721a9c69fe4c33cb9Atom transfer radical polymerization of dimethyl(methacryloyloxymethyl) phosphonateMukumoto, Kosuke; Zhong, Mingjiang; Matyjaszewski, KrzysztofEuropean Polymer Journal (2014), 56 (), 11-16CODEN: EUPJAG; ISSN:0014-3057. (Elsevier Ltd.)Atom transfer radical polymn. (ATRP) of dimethyl(methacryloyloxymethyl) phosphonate (DMMAMP, also referred to as MAPC1) with a copper/2,2'-bipyridine based catalytic system in DMSO resulted in a homopolymer with narrow mol. wt. distribution. Initiators for continuous activator regeneration (ICAR) ATRP of this monomer was also successful when using tris[(2-pyridyl)methyl]amine as ligand and azobisisobutyronitrile as supplementary initiator for the regeneration of the CuI activator. Normal ATRP provided better control compared to ICAR ATRP. Differential scanning calorimetry measurements of the resulting poly(DMMAMP) revealed a glass transition temp. (Tg = 63 °C), lower than that of poly(Me methacrylate). Chain extension of the poly(dimethyl(methacryloyloxymethyl) phosphonate) macroinitiator with styrene was successfully achieved with a high fraction of block copolymer.
- 35Charleux, B.; Delaittre, G.; Rieger, J.; D’Agosto, F. Macromolecules 2012, 45, 6753– 6765 DOI: 10.1021/ma300713fThere is no corresponding record for this reference.
- 36Warren, N. J.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 10174– 10185 DOI: 10.1021/ja502843f36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVGlu7bM&md5=a8ab4eb77ba9f218d45dc9cbb4e9fbf7Polymerization-Induced Self-Assembly of Block Copolymer Nano-objects via RAFT Aqueous Dispersion PolymerizationWarren, Nicholas J.; Armes, Steven P.Journal of the American Chemical Society (2014), 136 (29), 10174-10185CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A review. In this Perspective, we discuss the recent development of polymn.-induced self-assembly mediated by reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. This approach has quickly become a powerful and versatile technique for the synthesis of a wide range of bespoke org. diblock copolymer nano-objects of controllable size, morphol., and surface functionality. Given its potential scalability, such environmentally-friendly formulations are expected to offer many potential applications, such as novel Pickering emulsifiers, efficient microencapsulation vehicles, and sterilizable thermo-responsive hydrogels for the cost-effective long-term storage of mammalian cells.
- 37Sun, J.-T.; Hong, C.-Y.; Pan, C.-Y. Polym. Chem. 2013, 4, 873– 881 DOI: 10.1039/C2PY20612A37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlelt7g%253D&md5=6d4686d05496bb207018c7ec714735fdRecent advances in RAFT dispersion polymerization for preparation of block copolymer aggregatesSun, Jiao-Tong; Hong, Chun-Yan; Pan, Cai-YuanPolymer Chemistry (2013), 4 (4), 873-881CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A review. Differently from bulk, soln., suspension, emulsion, and miniemulsion polymns., the controlled radical dispersion polymn. (CRDP) demonstrates self-assembly of the block copolymers formed in the homogeneous system, forming various kinds of micelles or vesicles. Thus, this technol. can prep. both the block copolymers and the polymeric aggregates directly. Among CRDP, the reversible addn.-fragmentation chain transfer (RAFT) dispersion polymn. has been studied in relative detail and has been successfully developed to prep. a diverse range of assemblies. Several typical systems for RAFT dispersion polymn. are presented in detail and the factors influencing the polymn. and the in situ self-assembly are also highlighted in this minireview.
- 38Cunningham, V. J.; Alswieleh, A. M.; Thompson, K. L.; Williams, M.; Leggett, G. J.; Armes, S. P.; Musa, O. M. Macromolecules 2014, 47, 5613– 5623 DOI: 10.1021/ma501140hThere is no corresponding record for this reference.
- 39Blanazs, A.; Ryan, A. J.; Armes, S. P. Macromolecules 2012, 45, 5099– 5107 DOI: 10.1021/ma301059rThere is no corresponding record for this reference.
- 40An, Z.; Shi, Q.; Tang, W.; Tsung, C.-K.; Hawker, C. J.; Stucky, G. D. J. Am. Chem. Soc. 2007, 129, 14493– 14499 DOI: 10.1021/ja0756974There is no corresponding record for this reference.
- 41Rieger, J.; Grazon, C.; Charleux, B.; Alaimo, D.; Jérôme, C. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 2373– 2390 DOI: 10.1002/pola.2332941https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXltVOjsb4%253D&md5=216bf726a4cfe91f8e0935c65fd8d625Pegylated thermally responsive block copolymer micelles and nanogels via in situ RAFT aqueous dispersion polymerizationRieger, Jutta; Grazon, Chloe; Charleux, Bernadette; Alaimo, David; Jerome, ChristineJournal of Polymer Science, Part A: Polymer Chemistry (2009), 47 (9), 2373-2390CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A straightforward approach was developed to synthesize pegylated thermo-responsive core-shell nanoparticles in a min. of steps, directly in water. It is based on RAFT-controlled radical crosslinking copolymn. of N,N-diethylacrylamide (DEAAm) and N,N'-methylene bisacrylamide (MBA) in aq. dispersion polymn. Because DEAAm is water-sol. and poly(N,N-diethylacrylamide) (PDEAAm) exhibits a lower crit. soln. temp. at 32°, the initial medium was homogeneous, whereas the polymer formed a sep. phase at the reaction temp. The first macroRAFT agent was a surface-active trithiocarbonate based on a hydrophilic poly(ethylene oxide) block and a hydrophobic dodecyl chain. It was further extended with N,N-dimethylacrylamide (DMAAm) to target macroRAFT agents with increasing chain length. All macroRAFT agents provided excellent control over the aq. dispersion homopolymn. of DEAAm. When they were used in the radical crosslinking copolymn. of DEAAm and MBA, the stability and size of the resulting gel particles were found to depend strongly on the chain length of the macroRAFT agent, on the concns. of both the monomer and the crosslinker, and on the process (one step or two steps). The best-suited exptl. conditions to reach thermosensitive hydrogels with nanometric size and well-defined surface properties were detd.
- 42Li, Y.; Armes, S. P. Angew. Chem., Int. Ed. 2010, 49, 4042– 4046 DOI: 10.1002/anie.201001461There is no corresponding record for this reference.
- 43Sugihara, S.; Blanazs, A.; Armes, S. P.; Ryan, A. J.; Lewis, A. L. J. Am. Chem. Soc. 2011, 133, 15707– 15713 DOI: 10.1021/ja205887vThere is no corresponding record for this reference.
- 44Liu, G.; Qiu, Q.; Shen, W.; An, Z. Macromolecules 2011, 44, 5237– 5245 DOI: 10.1021/ma200984hThere is no corresponding record for this reference.
- 45Chambon, P.; Blanazs, A.; Battaglia, G.; Armes, S. P. Macromolecules 2012, 45, 5081– 5090 DOI: 10.1021/ma300816mThere is no corresponding record for this reference.
- 46Ladmiral, V.; Semsarilar, M.; Canton, I.; Armes, S. P. J. Am. Chem. Soc. 2013, 135, 13574– 13581 DOI: 10.1021/ja407033xThere is no corresponding record for this reference.
- 47Figg, C. A.; Simula, A.; Gebre, K. A.; Tucker, B. S.; Haddleton, D.; Sumerlin, B. S. Chem. Sci. 2015, 6, 1230 DOI: 10.1039/C4SC03334E47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvFOitLzI&md5=ddf49e8d73521f3e1d95d304f275f33cPolymerization-induced thermal self-assembly (PITSA)Figg, C. Adrian; Simula, Alexandre; Gebre, Kalkidan A.; Tucker, Bryan S.; Haddleton, David M.; Sumerlin, Brent S.Chemical Science (2015), 6 (2), 1230-1236CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Polymn.-induced self-assembly (PISA) is a versatile technique to achieve a wide range of polymeric nanoparticle morphologies. Most previous examples of self-assembled soft nanoparticle synthesis by PISA rely on a growing solvophobic polymer block that leads to changes in nanoparticle architecture during polymn. in a selective solvent. However, synthesis of block copolymers with a growing stimuli-responsive block to form various nanoparticle shapes has yet to be reported. This new concept using thermo-responsive polymers is termed polymn.-induced thermal self-assembly (PITSA). A reversible addn.-fragmentation chain transfer (RAFT) polymn. of N-isopropylacrylamide from a hydrophilic chain transfer agent composed of N,N-dimethylacrylamide and acrylic acid was carried out in water above the known lower crit. soln. temp. (LCST) of poly(N-isopropylacrylamide) (PNIPAm). After reaching a certain chain length, the growing PNIPAm self-assembled, as induced by the LCST, into block copolymer aggregates within which dispersion polymn. continued. To characterize the nanoparticles at ambient temps. without their dissoln., the particles were crosslinked immediately following polymn. at elevated temps. via the reaction of the acid groups with a diamine in the presence of a carbodiimide. Size exclusion chromatog. was used to evaluate the unimer mol. wt. distributions and reaction kinetics. Dynamic light scattering and transmission electron microscopy provided insight into the size and morphologies of the nanoparticles. The resulting block copolymers formed polymeric nanoparticles with a range of morphologies (e.g., micelles, worms, and vesicles), which were a function of the PNIPAm block length.
- 48Wan, W.-M.; Hong, C.-Y.; Pan, C.-Y. Chem. Commun. 2009, 5883– 5885 DOI: 10.1039/b912804bThere is no corresponding record for this reference.
- 49Semsarilar, M.; Jones, E. R.; Blanazs, A.; Armes, S. P. Adv. Mater. 2012, 24, 3378– 3382 DOI: 10.1002/adma.201200925There is no corresponding record for this reference.
- 50Zehm, D.; Ratcliffe, L. P. D.; Armes, S. P. Macromolecules 2013, 46, 128– 139 DOI: 10.1021/ma301459yThere is no corresponding record for this reference.
- 51Karagoz, B.; Boyer, C.; Davis, T. P. Macromol. Rapid Commun. 2014, 35, 417– 421 DOI: 10.1002/marc.201300730There is no corresponding record for this reference.
- 52Pei, Y.; Dharsana, N. C.; van Hensbergen, J. A.; Burford, R. P.; Roth, P. J.; Lowe, A. B. Soft Matter 2014, 10, 5787– 5796 DOI: 10.1039/C4SM00729HThere is no corresponding record for this reference.
- 53Zhang, W.-J.; Hong, C.-Y.; Pan, C.-Y. Macromolecules 2014, 47, 1664– 1671 DOI: 10.1021/ma402497yThere is no corresponding record for this reference.
- 54Bleach, R.; Karagoz, B.; Prakash, S. M.; Davis, T. P.; Boyer, C. ACS Macro Lett. 2014, 3, 591– 596 DOI: 10.1021/mz500195u54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpsFaqs78%253D&md5=ae73c889db0575c204c802e467a5a3a0In Situ Formation of Polymer-Gold Composite Nanoparticles with Tunable MorphologiesBleach, Richard; Karagoz, Bunyamin; Prakash, Shyam M.; Davis, Thomas P.; Boyer, CyrilleACS Macro Letters (2014), 3 (7), 591-596CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)A simple and efficient route to gold-polymer nanoparticle composites is described. Our versatile synthetic route exerts facile control over polymer nanoparticle morphol., including micelles, rod-like structures, and vesicles, all easily attainable from a single polymn. taken to different monomer conversions. Specifically, poly[oligo(ethylene glycol) methacrylate]-b-poly(dimethylaminoethyl methacrylate)-b-poly(styrene) (POEGMA-b-PDMAEMA-b-PST) triblock copolymers were synthesized using a polymn. induced self-assembly (PISA) approach. Subsequently, spherical gold nanoparticles (10 nm AuNPs) were formed at the hydrophilic-hydrophobic nexus of the assembled triblock copolymer nanoaggregates by the addn. of chloroauric acid (HAuCl4) followed by in situ redn. using NaBH4. After redn., the cloudy white nanoparticle dispersions turned to a red-purple color. The gold nanoparticles that formed were stabilized by the enveloping polymeric nanostructures, neither pptn. nor agglomeration occurred. We demonstrated that we were able to tune the gold nanoparticle compn. in these polymer-gold composites by varying the concn. of chloroauric acid. Morphol., particle size, mol. wt., AuNP content, and chem. structure of the polymer structures were characterized by transmittance electron microscopy (TEM), dynamic light scattering (DLS), size exclusion chromatog. (SEC), thermal gravimetric anal. (TGA), and 1H NMR. Finally, the formation of the AuNPs occurred without affecting the polymer nanoparticle morphol.
- 55Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Polym. Chem. 2014, 5, 3466– 3475 DOI: 10.1039/c4py00201f55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmtlGgsr8%253D&md5=5cc9158e0615ba6fe5993d8292ed4880Poly(methacrylic acid)-based AB and ABC block copolymer nano-objects prepared via RAFT alcoholic dispersion polymerizationSemsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P.Polymer Chemistry (2014), 5 (10), 3466-3475CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A series of well-defined amphiphilic poly(methacrylic acid)-poly(benzyl methacrylate) (PMAA-PBzMA) diblock copolymers were synthesized via polymn.-induced self-assembly using an alc. dispersion polymn. formulation. Chain growth was mediated via reversible addn.-fragmentation chain transfer polymn. (RAFT) chem. using a trithiocarbonate-based chain transfer agent (CTA) at 70 °C. The poly(methacrylic acid) block was sol. in ethanol and acts as a steric stabilizer for the growing insol. PBzMA chains, resulting in the in situ generation of diblock copolymer nano-objects in the form of spheres, worms or vesicles, depending on the precise reaction conditions. Copolymer morphologies can be covalently stabilized via crosslinking to prevent their dissocn. when transferred into aq. soln., which leads to the formation of highly anionic nano-objects due to ionization of the PMAA stabilizer chains. ABC triblock copolymer nanoparticles can also be prepd. using this approach, where the third block was based on the semi-fluorinated monomer, 2,2,2-trifluoroethyl methacrylate (TFEMA). GPC studies confirm that chain extension was efficient and high TFEMA conversions can be achieved. Microphase sepn. between the mutually incompatible PBzMA and semi-fluorinated PTFEMA core-forming blocks occurs, producing a range of remarkably complex semi-fluorinated triblock copolymer morphologies.
- 56Gonzato, C.; Semsarilar, M.; Jones, E. R.; Li, F.; Krooshof, G. J. P.; Wyman, P.; Mykhaylyk, O. O.; Tuinier, R.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 11100– 11106 DOI: 10.1021/ja505406sThere is no corresponding record for this reference.
- 57Houillot, L.; Bui, C.; Farcet, C. l.; Moire, C.; Raust, J.-A.; Pasch, H.; Save, M.; Charleux, B. ACS Appl. Mater. Interfaces 2010, 2, 434– 442 DOI: 10.1021/am900695aThere is no corresponding record for this reference.
- 58Deng, Y.; Yang, C.; Yuan, C.; Xu, Y.; Bernard, J.; Dai, L.; Gérard, J.-F. J. Polym. Sci., Part A: Polym. Chem. 2013, 51, 4558– 4564 DOI: 10.1002/pola.2687258https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtF2iu7zN&md5=222a788bfc01b859b7cbb6d15ef1f536Hybrid organic-inorganic block copolymer nano-objects from RAFT polymerization-induced self-assemblyDeng, Yuanming; Yang, Cangjie; Yuan, Conghui; Xu, Yiting; Bernard, Julien; Dai, Lizong; Gerard, Jean-FrancoisJournal of Polymer Science, Part A: Polymer Chemistry (2013), 51 (21), 4558-4564CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A simple route to org.-inorg. (O/I) nano-objects with different morphologies through polymn.-induced block copolymer self-assembly is described. The synthetic strategy relies on the chain-extension of polyhedral oligomeric silsesquioxanes (POSS)-contg. macro-CTA (PMAiBuPOSS13 and PMAiBuPOSS19) with styrene at 120 °C in octane, a selective solvent of the POSS-contg. block. The polymn. system was proven to afford a plethora of O/I nano-objects, such as spherical micelles, cylindrical micelles, and vesicles depending on the resp. molar masses of the PMAiBuPOSS and polystyrene (PS) blocks. The cooling procedure was also proven to be a crucial step to generate particles with a unique morphol. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013.
- 59Fielding, L. A.; Derry, M. J.; Ladmiral, V.; Rosselgong, J.; Rodrigues, A. M.; Ratcliffe, L. P. D.; Sugihara, S.; Armes, S. P. Chem. Sci. 2013, 4, 2081– 2087 DOI: 10.1039/c3sc50305d59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXltVOltrs%253D&md5=03d448f80f5c6e8a2ed9acf3b6fc8abbRAFT dispersion polymerization in non-polar solvents: facile production of block copolymer spheres, worms and vesicles in n-alkanesFielding, Lee A.; Derry, Matthew J.; Ladmiral, Vincent; Rosselgong, Julien; Rodrigues, Aurelie M.; Ratcliffe, Liam P. D.; Sugihara, Shinji; Armes, Steven P.Chemical Science (2013), 4 (5), 2081-2087CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Well-defined poly(lauryl methacrylate-benzyl methacrylate) (PLMA-PBzMA) diblock copolymer nanoparticles are prepd. in n-heptane at 90° via reversible addn.-fragmentation chain transfer (RAFT) polymn. Under these conditions, the PLMA macromol. chain transfer agent (macro-CTA) is sol. in n-heptane, whereas the growing PBzMA block quickly becomes insol. Thus this dispersion polymn. formulation leads to polymn.-induced self-assembly (PISA). Using a relatively long PLMA macro-CTA with a mean d.p. (DP) of 37 or higher leads to the formation of well-defined spherical nanoparticles of 41 to 139 nm diam., depending on the DP targeted for the PBzMA block. In contrast, TEM studies confirm that using a relatively short PLMA macro-CTA (DP = 17) enables both worm-like and vesicular morphologies to be produced, in addn. to the spherical phase. A detailed phase diagram has been elucidated for this more asym. diblock copolymer formulation, which ensures that each pure phase can be targeted reproducibly. 1H NMR spectroscopy confirmed that high BzMA monomer conversions (>97%) were achieved within 5 h, while GPC studies indicated that reasonably good blocking efficiencies and relatively low diblock copolymer polydispersities (Mw/Mn < 1.30) were obtained in most cases. Compared to prior literature reports, this all-methacrylic PISA formulation is particularly novel because: (i) it is the first time that higher order morphologies (e.g. worms and vesicles) have been accessed in non-polar solvents and (ii) such diblock copolymer nano-objects are expected to have potential boundary lubrication applications for engine oils.
- 60Dan, M.; Huo, F.; Zhang, X.; Wang, X.; Zhang, W. J. Polym. Sci., Part A: Polym. Chem. 2013, 51, 1573– 1584 DOI: 10.1002/pola.2652760https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtFSlsg%253D%253D&md5=e7d3dcc4b664e459d97fd5b1321b1ab5Dispersion RAFT polymerization of 4-vinylpyridine in toluene mediated with the macro-RAFT agent of polystyrene dithiobenzoate: Effect of the macro-RAFT agent chain length and growth of the block copolymer nano-objectsDan, Meihan; Huo, Fei; Zhang, Xu; Wang, Xiaohui; Zhang, WangqingJournal of Polymer Science, Part A: Polymer Chemistry (2013), 51 (7), 1573-1584CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)The dispersion reversible addn.-fragmentation chain transfer (RAFT) polymn. of 4-vinylpyridine in toluene in the presence of the polystyrene dithiobenzoate (PS-CTA) macro-RAFT agent with different chain length is discussed. The RAFT polymn. undergoes an initial slow homogeneous polymn. and a subsequent fast heterogeneous one. The RAFT polymn. rate is dependent on the PS-CTA chain length, and short PS-CTA generally leads to fast RAFT polymn. The dispersion RAFT polymn. induces the self-assembly of the in situ synthesized polystyrene-b-poly(4-vinylpyridine) block copolymer into highly concd. block copolymer nano-objects. The PS-CTA chain length exerts great influence on the particle nucleation and the size and morphol. of the block copolymer nano-objects. It is found, short PS-CTA leads to fast particle nucleation and tends to produce large-sized vesicles or large-compd. micelles, and long PS-CTA leads to formation of small-sized nanospheres. Comparison between the polymn.-induced self-assembly and self-assembly of block copolymer in the block-selective solvent is made, and the great difference between the two methods is demonstrated. The present study is anticipated to be useful to reveal the chain extension and the particle growth of block copolymer during the RAFT polymn. under dispersion condition. © 2013 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013.
- 61Tan, J.; Rao, X.; Yang, J.; Zeng, Z. Macromolecules 2013, 46, 8441– 8448 DOI: 10.1021/ma401909aThere is no corresponding record for this reference.
- 62Shi, P.; Li, Q.; He, X.; Li, S.; Sun, P.; Zhang, W. Macromolecules 2014, 47, 7442– 7452 DOI: 10.1021/ma501598kThere is no corresponding record for this reference.
- 63Huang, C.-Q.; Pan, C.-Y. Polymer 2010, 51, 5115– 5121 DOI: 10.1016/j.polymer.2010.08.056There is no corresponding record for this reference.
- 64Warren, N. J.; Mykhaylyk, O. O.; Mahmood, D.; Ryan, A. J.; Armes, S. P. J. Am. Chem. Soc. 2014, 136, 1023– 1033 DOI: 10.1021/ja410593nThere is no corresponding record for this reference.
- 65Zhao, W.; Gody, G.; Dong, S.; Zetterlund, P. B.; Perrier, S. Polym. Chem. 2014, 5, 6990– 7003 DOI: 10.1039/C4PY00855C65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhsVKnt7vE&md5=3612b5577d91990be47b88c39e830b8eOptimization of the RAFT polymerization conditions for the in situ formation of nano-objects via dispersion polymerization in alcoholic mediumZhao, Wei; Gody, Guillaume; Dong, Siming; Zetterlund, Per B.; Perrier, SebastienPolymer Chemistry (2014), 5 (24), 6990-7003CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Hydrophilic polymer brushes based on poly(ethylene glycol) Me ether acrylate (P(PEGA454)) or poly(ethylene glycol) Me ether methacrylate (P(PEGMA475)), both having a trithiocarbonate end group, were prepd. in water-dioxane (9 : 1) at 44 °C via RAFT polymn. and subsequently used in RAFT dispersion polymn. of styrene in isopropanol at 90 °C. RAFT reaction conditions were first optimized to prep. P(PEGA454) and P(PEGMA475) macro-RAFT agents at high monomer conversions (>90%) and very low fraction of dead chains (<1%). Both polymer brushes allowed the prepn. of well-defined amphiphilic diblock copolymers (P(PEGA454)-b-PS and P(PEGMA475)-b-PS) which self-assemble in situ into nano-objects with various morphologies. Using relatively long chain P(PEGA454) or P(PEGMA475) macro-RAFT agents (DP ~ 75) leads to the formation of near uniform spherical nanoparticles with diams. ranging from 30 to 140 nm, depending on the targeted DP of the PS block. In contrast, TEM and DLS studies demonstrated that using a shorter P(PEGA454) or P(PEGMA475) macro-RAFT agent (DP ~ 20) enables the formation of worm-like micelles, vesicles and large compd. vesicle morphologies in addn. to spheres. Cryo-TEM was used to confirm polymn. induced morphol. transition, rather than morphologies obtained via self-assembly driven by selective solvent or solvent evapn. during the prepn. of samples for characterization.
- 66Sugihara, S.; Armes, S. P.; Blanazs, A.; Lewis, A. L. Soft Matter 2011, 7, 10787– 10793 DOI: 10.1039/c1sm06593aThere is no corresponding record for this reference.
- 67He, W.-D.; Sun, X.-L.; Wan, W.-M.; Pan, C.-Y. Macromolecules 2011, 44, 3358– 3365 DOI: 10.1021/ma2000674There is no corresponding record for this reference.
- 68Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Langmuir 2012, 28, 914– 922 DOI: 10.1021/la203991yThere is no corresponding record for this reference.
- 69Semsarilar, M.; Ladmiral, V.; Blanazs, A.; Armes, S. P. Langmuir 2013, 29, 7416– 7424 DOI: 10.1021/la304279yThere is no corresponding record for this reference.
- 70Rieger, J.; Stoffelbach, F.; Bui, C.; Alaimo, D.; Jérôme, C.; Charleux, B. Macromolecules 2008, 41, 4065– 4068 DOI: 10.1021/ma800544vThere is no corresponding record for this reference.
- 71Zhang, W.; D’Agosto, F.; Boyron, O.; Rieger, J.; Charleux, B. Macromolecules 2012, 45, 4075– 4084 DOI: 10.1021/ma300596fThere is no corresponding record for this reference.
- 72Rieger, J.; Zhang, W.; Stoffelbach, F. O.; Charleux, B. Macromolecules 2010, 43, 6302– 6310 DOI: 10.1021/ma1009269There is no corresponding record for this reference.
- 73Zhang, W.; D’Agosto, F.; Dugas, P.-Y.; Rieger, J.; Charleux, B. Polymer 2013, 54, 2011– 2019 DOI: 10.1016/j.polymer.2012.12.028There is no corresponding record for this reference.
- 74Ratcliffe, L. P. D.; Blanazs, A.; Williams, C. N.; Brown, S. L.; Armes, S. P. Polym. Chem. 2014, 5, 3643– 3655 DOI: 10.1039/c4py00203b74https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXns1Wju7c%253D&md5=6d5f403dc2c766c3024f7e2ad246ea65RAFT polymerization of hydroxy-functional methacrylic monomers under heterogeneous conditions: effect of varying the core-forming blockRatcliffe, L. P. D.; Blanazs, A.; Williams, C. N.; Brown, S. L.; Armes, S. P.Polymer Chemistry (2014), 5 (11), 3643-3655CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Statistical copolymn. of a 1:1 molar ratio of a water-miscible monomer (2-hydroxyethyl methacrylate, HEMA) with a water-immiscible monomer (4-hydroxybutyl methacrylate, HBMA) has been conducted in water via reversible addn.-fragmentation chain transfer (RAFT) polymn. using a water-sol. poly(glycerol monomethacrylate) macromol. chain transfer agent (PGMA macro-CTA). In principle, such a hybrid formulation might be expected to be intermediate between RAFT dispersion polymn. and RAFT emulsion polymn. Under such circumstances, it is of particular interest to examine whether both monomers are actually consumed and, if so, whether their rates of reaction are comparable. Given the water-soly. of both the PGMA macro-CTA and the free radical azo initiator, it is perhaps counter-intuitive that the water-immiscible HBMA is initially consumed significantly faster than the water-miscible HEMA, as judged by 1H NMR studies of this copolymn. However, both comonomers are eventually almost fully consumed at 70 °C. A detailed phase diagram has been constructed for this RAFT formulation that enables reproducible syntheses of various pure copolymer morphologies, including spheres, worms and vesicles. It is emphasized that utilizing a 1:1 HEMA/HBMA molar ratio produces a core-forming statistical copolymer block that is isomeric with the poly(2-hydroxypropyl methacrylate) (PHPMA) core-forming block previously synthesized via RAFT aq. dispersion polymn. (see A. Blanazs et al., Macromols., 2012, 45, 5099-5107). Hence it is rather remarkable that the thermo-responsive behavior of PGMA-P(HBMA-stat-HEMA) statistical block copolymer worm gels differs qual. from that exhibited by PGMA-PHPMA diblock copolymer worm gels.
- 75Binauld, S.; Delafresnaye, L.; Charleux, B.; D’Agosto, F.; Lansalot, M. Macromolecules 2014, 47, 3461– 3472 DOI: 10.1021/ma402549xThere is no corresponding record for this reference.
- 76Zhang, X.; Boisson, F.; Colombani, O.; Chassenieux, C.; Charleux, B. Macromolecules 2014, 47, 51– 60 DOI: 10.1021/ma402125rThere is no corresponding record for this reference.
- 77Carlsson, L.; Fall, A.; Chaduc, I.; Wagberg, L.; Charleux, B.; Malmstrom, E.; D’Agosto, F.; Lansalot, M.; Carlmark, A. Polym. Chem. 2014, 5, 6076– 6086 DOI: 10.1039/C4PY00675E77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVylsrvI&md5=2210816d8f6fc6c301ea6d3aeb9d3593Modification of cellulose model surfaces by cationic polymer latexes prepared by RAFT-mediated surfactant-free emulsion polymerizationCarlsson, Linn; Fall, Andreas; Chaduc, Isabelle; Waagberg, Lars; Charleux, Bernadette; Malmstroem, Eva; D'Agosto, Franck; Lansalot, Muriel; Carlmark, AnnaPolymer Chemistry (2014), 5 (20), 6076-6086CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)This paper presents the successful surface modification of a model cellulose substrate by the prepn. and subsequent phys. adsorption of cationic polymer latexes. The first part of the work introduces novel charged polymer nanoparticles constituted of amphiphilic block copolymers based on cationic P(DMAEMA-co-MAA) as the hydrophilic segment, and PMMA as the hydrophobic segment. First, RAFT polymn. of DMAEMA in water was performed at pH 7, below its pKa. The simultaneous hydrolysis of DMAEMA led to the formation of a statistical copolymer incorporating mainly protonated DMAEMA units and some deprotonated methacrylic acid units at pH 7. The following step was the RAFT-mediated surfactant-free emulsion polymn. of MMA using P(DMAEMA-co-MAA) as a hydrophilic macromol. RAFT agent. The formed amphiphilic block copolymers self-assembled into cationic latex nanoparticles by polymn.-induced self-assembly. The nanoparticles were found to increase in size with increasing molar mass of the hydrophobic block. The cationic latexes were subsequently adsorbed to cellulose model surfaces in a quartz crystal microbalance equipment with dissipation. The adsorbed amt., in mg m-2, increased with increasing size of the nanoparticles. This approach allows for phys. surface modification of cellulose, utilizing a water suspension of particles for which both the surface chem. and the surface structure can be altered in a well-defined way.
- 78Zhang, X.; Cardozo, A. F.; Chen, S.; Zhang, W.; Julcour, C.; Lansalot, M.; Blanco, J.-F.; Gayet, F.; Delmas, H.; Charleux, B.; Manoury, E.; D’Agosto, F.; Poli, R. Chem. - Eur. J. 2014, 20, 15505– 15517 DOI: 10.1002/chem.20140381978https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVyks77I&md5=51101379f694359686096b0fc5e880daCore-Shell Nanoreactors for Efficient Aqueous Biphasic CatalysisZhang, Xuewei; Cardozo, Andres F.; Chen, Si; Zhang, Wenjing; Julcour, Carine; Lansalot, Muriel; Blanco, Jean-Francois; Gayet, Florence; Delmas, Henri; Charleux, Bernadette; Manoury, Eric; D'Agosto, Franck; Poli, RinaldoChemistry - A European Journal (2014), 20 (47), 15505-15517CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)Water-borne phosphine-functionalized core-cross-linked micelles (CCM) consisting of a hydrophobic core and a hydrophilic shell were obtained as stable latexes by reversible addn.-fragmentation chain transfer (RAFT) in water in a one-pot, three-step process. Initial homogeneous aq.-phase copolymn. of methacrylic acid (MAA) and poly(ethylene oxide) Me ether methacrylate (PEOMA) is followed by copolymn. of styrene (S) and 4-diphenylphosphinostyrene (DPPS), yielding P(MAA-co-PEOMA)-b-P(S-co-DPPS) amphiphilic block copolymer micelles (M) by polymn.-induced self-assembly (PISA), and final micellar crosslinking with a mixt. of S and diethylene glycol dimethacrylate. The CCM were characterized by dynamic light scattering and NMR spectroscopy to evaluate size, dispersity, stability, and the swelling ability of various org. substrates. Coordination of [Rh(acac)(CO)2] (acac=acetylacetonate) to the core-confined phosphine groups was rapid and quant. The CCM and M latexes were then used, in combination with [Rh(acac)(CO)2], to catalyze the aq. biphasic hydroformylation of 1-octene, in which they showed high activity, recyclability, protection of the activated Rh center by the polymer scaffold, and low Rh leaching. The CCM latex gave slightly lower catalytic activity but significantly less Rh leaching than the M latex. A control expt. conducted in the presence of the sulfoxantphos ligand pointed to the action of the CCM as catalytic nanoreactors with substrate and product transport into and out of the polymer core, rather than as a surfactant in interfacial catalysis.
- 79Ihli, J.; Bots, P.; Kulak, A.; Benning, L. G.; Meldrum, F. C. Adv. Funct. Mater. 2013, 23, 1965– 1973 DOI: 10.1002/adfm.201201742There is no corresponding record for this reference.
- 80Jones, E. R.; Semsarilar, M.; Blanazs, A.; Armes, S. P. Macromolecules 2012, 45, 5091– 5098 DOI: 10.1021/ma300898eThere is no corresponding record for this reference.
- 81Blanazs, A.; Madsen, J.; Battaglia, G.; Ryan, A. J.; Armes, S. P. J. Am. Chem. Soc. 2011, 133, 16581– 16587 DOI: 10.1021/ja206301aThere is no corresponding record for this reference.
- 82Odian, G. Principles of Polymerization, 4th ed.; Wiley: 2004.There is no corresponding record for this reference.
- 83Warren, N. J.; Mykhaylyk, O. O.; Ryan, A. J.; Williams, M.; Doussineau, T.; Dugourd, P.; Antoine, R.; Portale, G.; Armes, S. P. J. Am. Chem. Soc. 2015, 137, 1929– 1937 DOI: 10.1021/ja511423mThere is no corresponding record for this reference.
- 84Atkins, P.; De Paula, J. Physical Chemistry; Oxford University Press: 2006.There is no corresponding record for this reference.
- 85Semsarilar, M.; Jones, E. R.; Armes, S. P. Polym. Chem. 2014, 5, 195– 203 DOI: 10.1039/C3PY01042B85https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvVKjtb7L&md5=281b363840e59c0e81811b7ab500d3beComparison of pseudo-living character of RAFT polymerizations conducted under homogeneous and heterogeneous conditionsSemsarilar, Mona; Jones, Elizabeth R.; Armes, Steven P.Polymer Chemistry (2014), 5 (1), 195-203CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)RAFT dispersion polymn. of 2,2,2-trifluoroethyl methacrylate (TFEMA) is conducted in ethanol at 70 °C using either poly(2-(dimethylamino)ethyl methacrylate) or poly(methacrylic acid) as a macromol. chain transfer agent. If the diblock copolymer nanoparticles are not too large, the small refractive index difference between the PTFEMA cores and ethanol leads to minimal light scattering. This enables the pseudo-living character of RAFT formulations conducted under soln. and dispersion polymn. conditions to be compared by monitoring the loss of RAFT chain-ends via UV-visible absorption spectroscopy. Significantly fewer chain-ends are lost during RAFT dispersion polymn., suggesting that such heterogeneous formulations have greater pseudo-living character. Moreover, 19F NMR spectroscopy provides the first direct exptl. evidence that RAFT dispersion polymn. proceeds via monomer-swollen block copolymer micelles. The relatively low refractive index of PTFEMA complicates GPC anal., leading to apparent contamination of the diblock copolymer and erroneously high polydispersities. However, this artifact can be cor. by deconvolution of the GPC curves, followed by their reconstruction using appropriate refractive indexes.
- 86Blanazs, A.; Armes, S. P.; Ryan, A. J. Macromol. Rapid Commun. 2009, 30, 267– 277 DOI: 10.1002/marc.200800713There is no corresponding record for this reference.
- 87Lowenstam, H. A.; Weiner, S. On Biomineralization; Oxford University Press: New York, 1989.There is no corresponding record for this reference.
- 88Meldrum, F. C.; Cölfen, H. Chem. Rev. 2008, 108, 4332– 4432 DOI: 10.1021/cr8002856There is no corresponding record for this reference.
- 89Berman, A.; Hanson, J.; Leiserowitz, L.; Koetzle, T. F.; Weiner, S.; Addadi, L. Science 1993, 259, 776– 779 DOI: 10.1126/science.259.5096.776There is no corresponding record for this reference.
- 90Kulak, A. N.; Iddon, P.; Li, Y.; Armes, S. P.; Cölfen, H.; Paris, O.; Wilson, R. M.; Meldrum, F. C. J. Am. Chem. Soc. 2007, 129, 3729– 3736 DOI: 10.1021/ja067422eThere is no corresponding record for this reference.
- 91Kim, Y.-Y.; Ganesan, K.; Yang, P.; Kulak, A. N.; Borukhin, S.; Pechook, S.; Ribeiro, L.; Kröger, R.; Eichhorn, S. J.; Armes, S. P.; Pokroy, B.; Meldrum, F. C. Nat. Mater. 2011, 10, 890– 896 DOI: 10.1038/nmat3103There is no corresponding record for this reference.
- 92Kulak, A. N.; Yang, P.; Kim, Y.-Y.; Armes, S. P.; Meldrum, F. C. Chem. Commun. 2014, 50, 67– 69 DOI: 10.1039/C3CC47904HThere is no corresponding record for this reference.
- 93Kulak, A. N.; Semsarilar, M.; Kim, Y.-Y.; Ihli, J.; Fielding, L. A.; Cespedes, O.; Armes, S. P.; Meldrum, F. C. Chem. Sci. 2014, 5, 738– 743 DOI: 10.1039/C3SC52615A93https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXitVSqtb7J&md5=71fa681a2fb3d6f8c55fa12f7f6de82bOne-pot synthesis of an inorganic heterostructure: uniform occlusion of magnetite nanoparticles within calcite single crystalsKulak, Alexander N.; Semsarilar, Mona; Kim, Yi-Yeoun; Ihli, Johannes; Fielding, Lee A.; Cespedes, Oscar; Armes, Steven P.; Meldrum, Fiona C.Chemical Science (2014), 5 (2), 738-743CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A facile one-pot method is described for the formation of novel heterostructures in which inorg. nanoparticles are homogeneously distributed throughout an inorg. single crystal matrix. Our strategy uses nanoparticles functionalised with a poly(sodium 4-styrenesulfonate)-poly(methacrylic acid) [PNaStS-PMAA] diblock copolymer as a sol. crystal growth additive. This copolymer plays a no. of essential roles. The PMAA anchor block is phys. adsorbed onto the inorg. nanoparticles, while the PNaStS block acts as an electrosteric stabilizer and ensures that the nanoparticles retain their colloidal stability in the crystal growth soln. In addn., this strong acid block promotes binding to both the nanoparticles and the host crystal, which controls nanoparticle incorporation within the host crystal lattice. We show that this approach can be used to achieve encapsulation loadings of at least 12 wt% copolymer-coated magnetite particles within calcite single crystals. Transmission electron microscopy shows that these nanoparticles are uniformly distributed throughout the calcite, and that the crystal lattice retains its continuity around the embedded magnetite particles. Characterization of these calcite/magnetite nanocomposites confirmed their magnetic properties. This new exptl. approach is expected to be quite general, such that a small family of block copolymers could be used to drive the incorporation of a wide range of pre-prepd. nanoparticles into host crystals, giving intimate mixing of phases with contrasting properties, while limiting nanoparticle aggregation and migration.
- 94Borukhin, S.; Bloch, L.; Radlauer, T.; Hill, A. H.; Fitch, A. N.; Pokroy, B. Adv. Funct. Mater. 2012, 22, 4216– 4224 DOI: 10.1002/adfm.201201079There is no corresponding record for this reference.
- 95Kim, Y.-Y.; Ribeiro, L.; Maillot, F.; Ward, O.; Eichhorn, S. J.; Meldrum, F. C. Adv. Mater. 2010, 22, 2082– 2086 DOI: 10.1002/adma.200903743There is no corresponding record for this reference.
- 96Tomson, M. B. J. Cryst. Growth 1983, 62, 106– 112 DOI: 10.1016/0022-0248(83)90013-196https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksVehtLk%253D&md5=a1801446c6161015b96f941e19a41b0dEffect of precipitation inhibitors on calcium carbonate scale formationTomson, Mason B.Journal of Crystal Growth (1983), 62 (1), 106-12CODEN: JCRGAE; ISSN:0022-0248.A model of threshold scale inhibition is proposed based on the relative diffusion rates of the CO32- and the inhibitor ions, and a criterion for the min. concn. of inhibitor, is developed and tested. The relation ZIn(In2-) > 2(CO32-), [where ZIn is the magnitude of the charge on the inhibitor ion and () denotes molar concn.] gave good agreement with exptl. values.
- 97Ketrane, R.; Saidani, B.; Gil, O.; Leleyter, L.; Baraud, F. Desalination 2009, 249, 1397– 1404 DOI: 10.1016/j.desal.2009.06.013There is no corresponding record for this reference.
- 98Ning, Y.; Fielding, L. A.; Andrews, T. S.; Growney, D. J.; Armes, S. P. Sulfate-based anionic diblock copolymer nanoparticles for efficient occlusion within zinc oxide Nanoscale 2015, 7, 6691– 6702There is no corresponding record for this reference.
Supporting Information
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.5b02212.
31P NMR spectra and GPC data for the three PMPA macro-CTAs, GPC and DLS data for kinetics of dispersion polymerization experiments, GPC data for all PMPx–PBzMAy diblock copolymers, additional TEM images, DLS data for aqueous dispersion polymerizations conducted using binary mixtures of macro-CTAs, and Raman spectra for selected CaCO3 occlusion experiments (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.