Solution and Solid-State Behavior of Amphiphilic ABA Triblock Copolymers of Poly(acrylic acid-stat-styrene)-block-poly(butyl acrylate)-block-poly(acrylic acid-stat-styrene)Click to copy article linkArticle link copied!
- Thomas J. NealThomas J. NealDepartment of Chemistry, The University of Sheffield, Dainton Building, Sheffield, South YourkshireS3 7HF, U.K.More by Thomas J. Neal
- Robert D. BradleyRobert D. BradleyAkzoNobel Decorative Paints, Wexham Road, Slough, BerkshireSL2 5DS, U.K.More by Robert D. Bradley
- Martin W. MurrayMartin W. MurrayAkzoNobel Decorative Paints, Wexham Road, Slough, BerkshireSL2 5DS, U.K.More by Martin W. Murray
- Neal S. J. WilliamsNeal S. J. WilliamsAkzoNobel Decorative Paints, Wexham Road, Slough, BerkshireSL2 5DS, U.K.More by Neal S. J. Williams
- Simon N. EmmettSimon N. EmmettAkzoNobel Decorative Paints, Wexham Road, Slough, BerkshireSL2 5DS, U.K.More by Simon N. Emmett
- Anthony J. RyanAnthony J. RyanDepartment of Chemistry, The University of Sheffield, Dainton Building, Sheffield, South YourkshireS3 7HF, U.K.More by Anthony J. Ryan
- Sebastian G. Spain*Sebastian G. Spain*Email: [email protected]Department of Chemistry, The University of Sheffield, Dainton Building, Sheffield, South YourkshireS3 7HF, U.K.More by Sebastian G. Spain
- Oleksandr O. Mykhaylyk*Oleksandr O. Mykhaylyk*Email: [email protected]Department of Chemistry, The University of Sheffield, Dainton Building, Sheffield, South YourkshireS3 7HF, U.K.More by Oleksandr O. Mykhaylyk
Abstract
A combination of statistical and triblock copolymer properties is explored to produce stable aqueous polymer dispersions suitable for the film formation. In order to perform an extensive structural characterization of the products in the dissolved, dispersed, and solid states, a wide range of symmetrical poly(acrylic acid-stat-styrene)x-block-poly(butyl acrylate)y-block-poly(acrylic acid-stat-styrene)x, poly(AA-st-St)x-b-PBAy-b-poly(AA-st-St)x, (x = 56, 108 and 140, y = 100–750; the AA:St molar ratio is 42:58) triblock copolymers were synthesized by reversible addition–fragmentation chain transfer (RAFT) solution polymerization using a bifunctional symmetrical RAFT agent. It is demonstrated that the amphiphilic statistical outer blocks can provide sufficient stabilization to largely hydrophobic particles in aqueous dispersions. Such a molecular design provides an advantage over copolymers composed only of homoblocks, as a simple variation of the statistical block component ratio provides an efficient way to control the hydrophilicity of the stabilizer block, which ultimately affects the copolymer morphology in solutions and solid films. It was found by small-angle X-ray scattering (SAXS) that the copolymers behaved as dissolved chains in methylethylketone (MEK) but self-assembled in water into stable and well-defined spherical particles that increased in size with the length of the hydrophobic PBA block. These particles possessed an additional particulate surface structure formed by the statistical copolymer stabilizer block, which self-folded through the hydrophobic interactions between the styrene units. SAXS and atomic force microscopy showed that the copolymer films cast from the MEK solutions formed structures predicted by self-consistent field theory for symmetrical triblock copolymers, while the aqueous dispersions formed structural morphologies similar to a close-packed spheres, as would be expected for copolymer particles trapped kinetically due to the restricted movement of the blocks in the initial aqueous dispersion. A strong correlation between the structural morphology and mechanical properties of the films was observed. It was found that the properties of the solvent cast films were highly dependent on the ratios of the hard [poly(AA-st-St)] and soft (PBA) blocks, while the aqueous cast films did not show such a dependence. The continuous phase of hard blocks, always formed in the case of the aqueous cast films, produced films with a higher elastic modulus and a lower extension-to-break in a comparison with the solvent-cast films.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Introduction
Experimental Section
Materials
Synthesis of P(AA-st-St) Macro-CTA via RAFT Solution Polymerization
NMR | SECc | |||||
---|---|---|---|---|---|---|
copolymer | conversion (%) | DP of copolymera | mole fraction of AAb | Mn (kDa) | Mw (kDa) | Mw/Mn |
A56A56 | 68 | 112 (56 + 56) | 0.42 | 9.1 | 11.4 | 1.26 |
A108A108 | 58 | 216 (108 + 108) | 0.42 | 15.9 | 19.8 | 1.24 |
A140A140 | 56 | 280 (140 + 140) | 0.42 | 19.2 | 24.5 | 1.28 |
The total DP of both A blocks used to synthesize the ABA triblock, where the A blocks will be half the DP of the respective macro-CTA. The DPs were calculated using the ratio of the integrals from the RAFT chain end and the copolymer backbone.
The copolymer composition was calculated by examining the respective integrals in the 1H NMR spectrum in DMSO.
SEC measurements were performed using a THF eluent containing 1% v/v acetic acid against PSt standards.
Parallel Syntheses of Triblock Copolymers via RAFT Solution Polymerization
1H Nuclear Magnetic Resonance (NMR) Spectroscopy
Size Exclusion Chromatography/Advanced Polymer Chromatography (SEC/APC)
Small-Angle X-ray Scattering (SAXS) Measurements
Grazing-Incident Small-Angle X-ray Scattering (GISAXS) Measurements
Atomic Force Microscopy (AFM)
Mechanical Testing
Results and Discussion
Synthesis and Characterization of Copolymers
NMR | SECa | ||||||
---|---|---|---|---|---|---|---|
macro-CTA | triblock | targeted DP (BA) | conversion (%) | synthesized DP (BA) | Mn (kDa) | Mw (kDa) | Mw/Mn |
A56A56 | A56B100A56 | 100 | 89 | 89 | 11400 | 15400 | 1.34 |
A56B150A56 | 150 | 93 | 140 | 13400 | 18100 | 1.35 | |
A56B200A56 | 200 | 92 | 184 | 14500 | 19700 | 1.25 | |
A56B300A56 | 300 | 92 | 276 | 18500 | 25000 | 1.35 | |
A56B500A56 | 500 | 92 | 460 | 24800 | 33100 | 1.34 | |
A56B750A56 | 750 | 88 | 660 | 32200 | 41500 | 1.29 | |
A108A108 | A108B100A108 | 100 | 78 | 78 | 17300 | 21600 | 1.25 |
A108B150A108 | 150 | 82 | 123 | 18000 | 23000 | 1.28 | |
A108B200A108 | 200 | 86 | 172 | 21300 | 26200 | 1.23 | |
A108B300A108 | 300 | 85 | 255 | 19897 | 26300 | 1.32 | |
A108B500A108 | 500 | 88 | 440 | 25300 | 35000 | 1.38 | |
A108B750A108 | 750 | 86 | 645 | 29300 | 42000 | 1.43 | |
A140A140 | A140B100A140 | 100 | 74 | 74 | 20500 | 25800 | 1.26 |
A140B150A140 | 150 | 72 | 108 | 22500 | 27900 | 1.24 | |
A140B200A140 | 200 | 71 | 142 | 22900 | 29100 | 1.27 | |
A140B300A140 | 300 | 76 | 228 | 24900 | 32300 | 1.30 | |
A140B500A140 | 500 | 72 | 360 | 25300 | 32200 | 1.28 | |
A140B750A140 | 750 | 75 | 563 | 34600 | 47200 | 1.37 |
SEC measurements were performed using a THF eluent containing 1% v/v acetic acid against PSt standards on an APC instrument.
Structure of the Macro-CTAs and Triblock Copolymers in MEK and Water Solutions
copolymer | Rcore (Å) | σcore (Å) | r2 (Å) |
---|---|---|---|
A56A56 | 16b | 5 | |
A108A108 | 16b | 7 | |
A140A140 | 17b | 7 | |
A56B100A56 | 53 | 14 | 12 |
A56B150A56 | 90 | 13 | 12 |
A56B200A56 | 119 | 18 | 12 |
A56B300A56 | 156 | 20 | 12 |
A56B500A56 | 247 | 28 | 12 |
A56B750A56 | 281 | 52 | 12 |
A108B100A108 | 18 | 9 | 15 |
A108B150A108 | 55 | 34 | 15 |
A108B200A108 | 99 | 24 | 15 |
A108B300A108 | 120 | 22 | 15 |
A108B500A108 | 191 | 39 | 15 |
A108B750A108 | 219 | 52 | 15 |
A140B100A140 | 15 | 5 | 16 |
A140B150A140 | 26 | 10 | 16 |
A140B200A140 | 32 | 23 | 16 |
A140B300A140 | 101 | 41 | 16 |
A140B500A140 | 150 | 41 | 16 |
A140B750A140 | 219 | 50 | 16 |
Rcore is the mean particle core radius, σcore is the standard deviation of the mean particle core radius, and r2 is the radius of the self-folded P(AA-st-St) chain on the surface of the particle (this value based upon the volume of the hard block is fixed throughout the fitting).
Note: Rcore refers to the total particle radius for the A56A56, A108A108, and A140A140 macro-CTAs.
Structural Characterization of Triblock Copolymer Films Cast from an Organic Solvent
Structural Characterization of Triblock Copolymer Films Cast from Aqueous Dispersions
Structural Characterization of Annealed Triblock Copolymer Films Cast from Aqueous Dispersion
Mechanical Characterization of the Triblock Copolymer Films
Conclusions
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.macromol.2c01299.
Further characterization of triblock copolymer solutions, dispersions films, and SAXS models (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.
Acknowledgments
AkzoNobel (Slough, UK) and EPSRC (EP/L016281/1) are thanked for funding a CDT Ph.D. CASE studentship of T.J.N. O.O.M. thanks EPSRC for the capital equipment grants to purchase (EP/M028437/1) and upgrade (EP/V034804/1) the laboratory-based Xenocs/Excillum SAXS instrument used to characterize the studied copolymer dispersions.
References
This article references 79 other publications.
- 1Bradford, E. B.; McKeever, L. D. Block Copolymers. Prog. Polym. Sci. 1971, 3, 109– 143, DOI: 10.1016/0079-6700(71)90003-7Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE38XhtVSjtb8%253D&md5=c0d01b0a8f622d4c9873aa0e5f40137dBlock copolymersBradford, E. B.; McKeever, L. D.Progress in Polymer Science (1971), 3 (), 109-43CODEN: PRPSB8; ISSN:0079-6700.A review with 234 refs. discusses the synthesis, characterization, properties, and uses of block copolymers.
- 2Bates, F. S.; Fredrickson, G. H. Block Copolymers-Designer Soft Materials. Phys. Today 1999, 52 (2), 32– 38, DOI: 10.1063/1.882522Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlymsr8%253D&md5=dd0e3bc303e671a3ae845969248fbc5eBlock copolymers-designer soft materialsBates, Frank S.; Fredrickson, Glenn H.Physics Today (1999), 52 (2), 32-38CODEN: PHTOAD; ISSN:0031-9228. (American Institute of Physics)A review with 18 refs. discusses the structure, morphol., phase behavior, and self-assembly properties of block copolymers.
- 3Amphiphilic Block Copolymers: Self-Assembly and Applications; Alexandridis, P.; Lindman, B., Eds.; Elsevier: Amsterdam, The Netherlands, 2000.Google ScholarThere is no corresponding record for this reference.
- 4Zhang, L.; Yu, K.; Eisenberg, A. Ion-Induced Morphological Changes in “Crew-Cut” Aggregates of Amphiphilic Block Copolymers. Science 1996, 272, 1777– 1779, DOI: 10.1126/science.272.5269.1777Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjslWnurk%253D&md5=9c5f9e9eaeb98318d1d10c2a812316d3Ion-induced morphological changes in "crew-cut" aggregates of amphiphilic block copolymersZhang, Lifeng; Yu, Kui; Eisenberg, AdiScience (Washington, D. C.) (1996), 272 (5269), 1777-1779CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The addn. of ions in micromolar (CaCl2 or HCl) or millimolar (NaCl) concns. can change the morphol. of "crew-cut" aggregates of amphiphilic block copolymers in dil. solns. In addn. to spherical, rodlike, and univesicular or lamellar aggregates, an unusual large compd. vesicle morphol. can be obtained from a single block copolymer. Some features of the spontaneously formed large compd. vesicles may make them esp. useful as vehicles for delivering drugs and as models of biol. cells. Gelation of a dil. spherical micelle soln. can also be induced by ions as the result of the formation of a cross-linked "pearl necklace" morphol.
- 5Zhang, L.; Eisenberg, A. Multiple Morphologies of “Crew-Cut” Aggregates of Polystyrene-b-Poly(Acrylic Acid) Block Copolymers. Science 1995, 268, 1728– 1731, DOI: 10.1126/science.268.5218.1728Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmsFCisbk%253D&md5=dce426eb449e334f10438dbcf2d5aac6Multiple morphologies of "crew-cut" aggregates of polystyrene-b-poly(acrylic acid) block copolymersZhang, Lifeng; Eisenberg, AdiScience (Washington, D. C.) (1995), 268 (5218), 1728-31CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The observation by transmission electron microscopy of six different stable aggregate morphologies is reported for the same family of highly asym. styrene-acrylic acid diblock copolymers prepd. in a low-mol.-wt. solvent system. Four of the morphologies consist of spheres, rods, lamellae, and vesicles in aq. soln., whereas the fifth consists of simple reverse micelle-like aggregates. The sixth consists of up to micrometer-size spheres in aq. soln. that have hydrophilic surfaces and are filled with the reverse micelle-like aggregates. In addn., a needle-like solid, which is highly birefringent, is obtained on drying of aq. solns. of the spherical micelles. This range of morphologies is believed to be unprecedented for a block copolymer system.
- 6van Hest, J. C. M.; Delnoye, D. A. P.; Baars, M. W. P. L.; van Genderen, M. H. P.; Meijer, E. W. Polystyrene-Dendrimer Amphiphilic Block Copolymers with a Generation-Dependent Aggregation. Science 1995, 268 (268), 1592– 1595, DOI: 10.1126/science.268.5217.1592Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmtlarsbo%253D&md5=b4d4f8c6bbb3715ad64c022131a70bd7Polystyrene-dendrimer amphiphilic block copolymers with a generation-dependent aggregationvan Hest, J. C. M.; Delnoye, D. A. P.; Baars, M. W. P. L.; van Genderen, M. H. P.; Meijer, E. W.Science (Washington, D. C.) (1995), 268 (5217), 1592-5CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)A class of amphiphilic macromols. has been synthesized by combining well-defined polystyrene (PS) with poly(propylene imine) dendrimers. Five different generations, from PS-dendr-NH2 up to Ps-dendr-(NH2)32, were prepd. in yields of 70 to 90 percent. Dynamic light scattering, cond. measurements, and transmission electron microscopy show that in aq. phases, PS-dendr-(NH2)32 forms spherical micelles, PS-dendr-(NH2)16 forms micellar rods, and PS-dendr-(NH2)8 forms vesicular structures. The lower generations of this class of macromols. show inverted micellar behavior. The obsd. effect of amphiphile geometry on aggregation behavior is in qual. agreement with the theory of J. N. Israelachvili, et al. (1976). The amphiphiles presented here are similar in shape but different in size as compared with traditional surfactants, whereas they are similar in size but different in shape as compared with traditional block copolymers.
- 7Bates, F. S. Polymer-Polymer Phase Behaviour. Science 1991, 251 (251), 898– 905, DOI: 10.1126/science.251.4996.898Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhsVCgsrk%253D&md5=53fc355f97cf74f3bf5e17e2c427509fPolymer-polymer phase behaviorBates, Frank S.Science (Washington, DC, United States) (1991), 251 (4996), 898-905CODEN: SCIEAS; ISSN:0036-8075.A review with 57 refs. on basic factors governing polymer-polymer phase behavior with emphasis on linear homopolymer blends and diblock copolymers as representative model mol. architectures. Equil. thermodn. and phase sepn. dynamics are reviewed and discussed.
- 8Mai, Y.; Eisenberg, A. Self-Assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41 (18), 5969– 5985, DOI: 10.1039/c2cs35115cGoogle Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1aqsbvL&md5=37964bf5011f9bc7a6aa42c4b612cf91Self-assembly of block copolymersMai, Yiyong; Eisenberg, AdiChemical Society Reviews (2012), 41 (18), 5969-5985CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Block copolymer (BCP) self-assembly has attracted considerable attention for many decades. The present tutorial review introduces the primary principles of BCP self-assembly in bulk and in soln., by describing expts., theories, accessible morphologies, etc.
- 9Xu, R.; Winnik, M. A.; Hallett, F. R.; Riess, G.; Croucher, M. D. Light-Scattering Study of the Association Behavior of Styrene-Ethylene Oxide Block Copolymers in Aqueous Solution. Macromolecules 1991, 24 (1), 87– 93, DOI: 10.1021/ma00001a014Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXisFSmsQ%253D%253D&md5=1f8d2b01b4aafa5d0b14de8711b0384dLight-scattering study of the association behavior of styrene-ethylene oxide block copolymers in aqueous solutionXu, Renliang; Winnik, Mitchell A.; Hallett, F. R.; Riess, Gerard; Croucher, Melvin D.Macromolecules (1991), 24 (1), 87-93CODEN: MAMOBX; ISSN:0024-9297.The micellization behavior of styrene (I)-ethylene oxide (II) diblock and triblock copolymers in aq. solns. is studied via static and dynamic light scattering. There are 2 narrowly distributed populations of particles in the soln. The sizes of both sets of particles are unchanged in concn. range of 2 × 10 -5 - 2 × 10-3 g/mL. The wt. fraction of the larger particles decreases with increasing polymer concn. The small particles are regular micelles with an assocn. no. of several hundred; the large particles are loose micellar clusters of tens of micelles. The micelle structure is core-shell with a core of I and a shell of II.
- 10Wilhelm, M.; Zhao, C. Le; Wang, Y.; Xu, R.; Winnik, M. A.; Mura, J. L.; Riess, G.; Croucher, M. D. Poly(Styrene-Ethylene Oxide) Block Copolymer Micelle Formation in Water: A Fluorescence Probe Study. Macromolecules 1991, 24 (5), 1033– 1040, DOI: 10.1021/ma00005a010Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhtVehtb8%253D&md5=5257329b56f4c306aa325112f726ae9fPoly(styrene-ethylene oxide) block copolymer micelle formation in water: a fluorescence probe studyWilhelm, Manfred; Zhao, Cheng Le; Wang, Yongcai; Xu, Renliang; Winnik, Mitchell A.; Mura, Jean Luc; Riess, Gerard; Croucher, Melvin D.Macromolecules (1991), 24 (5), 1033-40CODEN: MAMOBX; ISSN:0024-9297.Block copolymer micelle formation was studied by a combination of fluorescent and quasielastic light scattering (QELS) techniques. Ethylene oxide-styrene diblock and triblock copolymers, with mol. wts. 8500-29,000, formed spherical micelles in H2O over the entire concn. range over which QELS signals could be detected. Pyrene (I) in H2O (6 × 10-7 M) partitions between the aq. and micellar phases, accompanied by 3 changes in I spectroscopy. A red shift in the excitation spectrum, a change in the vibrational fine structure of I fluorescence, and an increase in the fluorescence decay time (from 200 to ∼350 ns) accompanied the transfer of I from an aq. to a hydrophobic micellar environment. From these data, crit. micelle concns. (1-5 mg/L) and partition coeffs. (3 × 105) were calcd.
- 11Prochazka, K.; Kiserow, D.; Ramireddy, C.; Tuzar, Z.; Munk, P.; Webber, S. E. Time-Resolved Fluorescence Studies of the Chain Dynamics of Naphthalene-Labeled Polystyrene-Block-Poly(Methacrylic Acid) Micelles in Aqueous Media. Macromolecules 1992, 25 (1), 454– 460, DOI: 10.1021/ma00027a071Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38Xkt1aqug%253D%253D&md5=dc8198b84ce62c1791e10510da215520Time-resolved fluorescence studies of the chain dynamics of naphthalene-labeled polystyrene-block-poly(methacrylic acid) micelles in aqueous mediaProchazka, K.; Kiserow, D.; Ramireddy, C.; Tuzar, Z.; Munk, P.; Webber, S. E.Macromolecules (1992), 25 (1), 454-60CODEN: MAMOBX; ISSN:0024-9297.A-B block copolymers of polystyrene and poly(methacrylic acid) and polystyrene and poly(tert-Bu methacrylate), both with short-chain oligovinyl-2-naphthalene moieties attached to the end of the polystyrene block, were prepd. by anionic polymn. and micelles were formed by hydrolysis. Intramol. excimer formation (which is controlled by the mobility of the pendant fluorescent groups and the polymer chain dynamics) was sterically hindered in the micellar cores as they become more compact. Both steady-state and time-resolved excimer fluorescence were significantly influenced by the gradual collapse of the micellar cores and the increase in segment d. within the cores with an increasing content of water. Changes in lifetimes and preexponential factors for naphthalene fluorescence (monomer as well as excimer) were sensitive indicators of micelle formation.
- 12Astafieva, I.; Zhong, X. F.; Eisenberg, A. Critical Micellization Phenomena in Block Polyelectrolyte Solutions. Macromolecules 1993, 26 (26), 7339– 7352, DOI: 10.1021/ma00078a034Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmslahur4%253D&md5=bd163a0318c8c488cb225d305f2ffc50Critical micellization phenomena in block polyelectrolyte solutionsAstafieva, Irina; Zhong, Xing Fu; Eisenberg, AdiMacromolecules (1993), 26 (26), 7339-52CODEN: MAMOBX; ISSN:0024-9297.Crit. micelle concn. (CMC) was measured for a range of block copolyelectrolytes based on styrene (the insol. block) and Na acrylate. The length of the styrene blocks was 6-110, while that of the polyelectrolyte was ∼300-1400. The fluorescent probe technique was used with pyrene as a probe mol., and the data were treated by 5 different methods. CMC results were interpolated for a const. polyelectrolyte block length of 1000 units. Changing the insol. block length from 6 to 110 lowed the CMC from 1.6 × 10-5 to 5 × 10-8 M. By contrast, changing the sol. block length from 300 to 1400 typically changed the CMC values by a factor of <2. For very short polystyrene block lengths, the CMC decreased very rapidly with increasing length of the insol. block. By contrast, for higher block lengths (>12) the drop in the CMC was much more gradual. The results are compared with those of previous studies of block copolymer micelles and theories of block copolymer micellization.
- 13Qin, A.; Tian, M.; Ramireddy, C.; Webber, S. E.; Munk, P.; Tuzar, Z. Polystyrene-Poly(Methacrylic Acid) Block Copolymer Micelles. Macromolecules 1994, 27 (1), 120– 126, DOI: 10.1021/ma00079a018Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXjtFOltw%253D%253D&md5=38e7c164b285a85c5e089d0da6b234aaPolystyrene-poly(methacrylic acid) block copolymer micellesQin, Anwei; Tian, Minmin; Ramireddy, C.; Webber, Stephen E.; Munk, Petr; Tuzar, ZdenekMacromolecules (1994), 27 (1), 120-6CODEN: MAMOBX; ISSN:0024-9297.Block copolymer micelles formed by diblock and triblock copolymers of styrene (I) and methacrylic acid (II) were characterized in soln. in a mixed solvent with 80 vol % of dioxane and 20 vol % of water. Methods of static light scattering, quasielastic light scattering, differential refractometry, viscometry, sedimentation velocity, and densitometry were used. No unattached unimer mols. were obsd. Three independent methods were employed for obtaining micellar wts. and agreed well with each other. No anomalous behavior was obsd. by any method. The micellar solns. contained almost exclusively single micelles; only a few samples (those producing the largest micelles) contained micellar clusters. The micelles behaved hydrodynamically and thermodynamically as impermeable spheres. The structure of the micellar shell was characterized in some detail. Relations between the aggregation no. and the hydrodynamic radius of the micelles on the one hand and the sizes of the I and II blocks on the other were presented as scaling-type phenomenol. equations.
- 14Guo, M.; Pitet, L. M.; Wyss, H. M.; Vos, M.; Dankers, P. Y. W.; Meijer, E. W. Tough Stimuli-Responsive Supramolecular Hydrogels with Hydrogen-Bonding Network Junctions. J. Am. Chem. Soc. 2014, 136 (19), 6969– 6977, DOI: 10.1021/ja500205vGoogle Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXns1CntLY%253D&md5=ad1f36275bc00ec6050c57e2e6973716Tough Stimuli-Responsive Supramolecular Hydrogels with Hydrogen-Bonding Network JunctionsGuo, Mingyu; Pitet, Louis M.; Wyss, Hans M.; Vos, Matthijn; Dankers, Patricia Y. W.; Meijer, E. W.Journal of the American Chemical Society (2014), 136 (19), 6969-6977CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Hydrogels were prepd. with phys. crosslinks comprising 2-ureido-4[1H]-pyrimidinone (UPy) hydrogen-bonding units within the backbone of segmented amphiphilic macromols. contg. hydrophilic poly(ethylene glycol) (PEG). The bulk materials adopt nanoscopic phys. crosslinks composed of UPy-UPy dimers embedded in segregated hydrophobic domains dispersed within the PEG matrix as confirmed by cryo-electron microscopy. The amphiphilic network was swollen with high wt. fractions of water (wH2O ≈ 0.8) owing to the high PEG wt. fraction within the pristine polymers (wPEG ≈ 0.9). Two different PEG chain lengths were investigated and illustrate the corresponding consequences of crosslink d. on mech. properties. The resulting hydrogels exhibited high strength and resilience upon deformation, consistent with a microphase sepd. network, in which the UPy-UPy interactions were adequately shielded within hydrophobic nanoscale pockets that maintain the network despite extensive water content. The cumulative result is a series of tough hydrogels with tunable mech. properties and tractable synthetic prepn. and processing. Furthermore, the melting transition of PEG in the dry polymer was shown to be an effective stimulus for shape memory behavior.
- 15Huo, M.; Yuan, J.; Tao, L.; Wei, Y. Redox-Responsive Polymers for Drug Delivery: From Molecular Design to Applications. Polym. Chem. 2014, 5 (5), 1519– 1528, DOI: 10.1039/C3PY01192EGoogle Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvFGhs74%253D&md5=87de58de0d857d687802195ecafd49c6Redox-responsive polymers for drug delivery: from molecular design to applicationsHuo, Meng; Yuan, Jinying; Tao, Lei; Wei, YenPolymer Chemistry (2014), 5 (5), 1519-1528CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A review. Glutathione has been regarded as a significant signal for distinguishing between tumor and normal tissue. Recently, reactive oxygen species have attracted much attention for their close connection with many diseases. Taking advantage of the physiol. signals, redox-responsive polymeric drug carriers constitute a significant research area in the various stimuli-responsive polymers for biomedical applications. During the rapid development of redox-responsive polymers, mol. design and related synthetic methodol. plays a crucial role. In this review, we discuss the redn.- and oxidn.-responsive polymeric drug carriers from the view of functional groups, as well as their applications in controlled release.
- 16Schattling, P.; Jochum, F. D.; Theato, P. Multi-Stimuli Responsive Polymers - the All-in-One Talents. Polym. Chem. 2014, 5 (1), 25– 36, DOI: 10.1039/C3PY00880KGoogle Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvVKjtLfJ&md5=9de59f5913441e9f2b3e7c77d803bb85Multi-stimuli responsive polymers - the all-in-one talentsSchattling, Philipp; Jochum, Florian D.; Theato, PatrickPolymer Chemistry (2014), 5 (1), 25-36CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Stimuli-responsive polymers have gained increasing attention, which is attributed to the manifold applications they can be used for. Several years' intensive research was invested in stimuli-responsive polymers. Their stimuli-responsiveness led not only to novel responsive groups, which enabled the translation of an external phys. impact into a change of a material property, but also to polymers that are equipped with more than one responsive group. The integration of several responsive moieties within one polymer yields smart polymers exhibiting complex responsive behavior of the polymers. This review summarises recent developments in the area of multi-stimuli responsive polymers, laying the focus on the improved, multifaceted response of polymer materials depending on the impact of several external stimuli.
- 17Sun, Y.; Wang, Z.; Li, Y.; Zhang, Z.; Zhang, W.; Pan, X.; Zhou, N.; Zhu, X. Photoresponsive Amphiphilic Macrocycles Containing Main-Chain Azobenzene Polymers. Macromol. Rapid Commun. 2015, 36, 1341– 1347, DOI: 10.1002/marc.201500136Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXnvFWmtr4%253D&md5=90de703b2124052f0c097b06a8bd09a0Photoresponsive Amphiphilic Macrocycles Containing Main-Chain Azobenzene PolymersSun, Yadong; Wang, Zhao; Li, Yiwen; Zhang, Zhengbiao; Zhang, Wei; Pan, Xiangqiang; Zhou, Nianchen; Zhu, XiulinMacromolecular Rapid Communications (2015), 36 (14), 1341-1347CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, the first example of photosensitive cyclic amphiphilic homopolymers consisting of multiple biphenyl azobenzene chromophores in the cyclic main chain tethered with hydrophilic tetraethylene glycol monomethyl ether units is presented. The synthetic approach involves sequentially performed thermal catalyzed "click" step-growth polymn. in bulk, and Cu(I)-catalyzed azide-alkyne cycloaddn. (CuAAC) intramol. cyclization from α-alkyne/ω-azide linear precursors. It is obsd. that such amphiphilic macrocycles exhibit increased glass transition temps. (Tg), slightly faster trans-cis-trans photoisomerization, and enhanced fluorescence emission intensity compared with the corresponding linear polymers. In addn., the cyclic amphiphilic homopolymers self-assemble into spherical nanoparticles with smaller sizes which possess slower photoresponsive behaviors in a tetrahydrofuran/water mixt. compared with those of the linear ones. All these interesting observations suggest that the cyclic topol. has a great influence on the phys. properties and self-assembly behavior of these photoresponsive amphiphilic macrocycles in general.
- 18Blanazs, A.; Madsen, J.; Battaglia, G.; Ryan, A. J.; Armes, S. P. Mechanistic Insights for Block Copolymer Morphologies: How Do Worms Form Vesicles?. J. Am. Chem. Soc. 2011, 133, 16581– 16587, DOI: 10.1021/ja206301aGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFCgsLfM&md5=2f1761acb3dac1734f68866820e00119Mechanistic Insights for Block Copolymer Morphologies: How Do Worms Form Vesicles?Blanazs, Adam; Madsen, Jeppe; Battaglia, Giuseppe; Ryan, Anthony J.; Armes, Steven P.Journal of the American Chemical Society (2011), 133 (41), 16581-16587CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Amphiphilic diblock copolymers composed of two covalently linked, chem. distinct chains can be considered to be biol. mimics of cell membrane-forming lipid mols., but with typically more than an order of magnitude increase in mol. wt. These macromol. amphiphiles are known to form a wide range of nanostructures (spheres, worms, vesicles, etc.) in solvents that are selective for one of the blocks. However, such self-assembly is usually limited to dil. copolymer solns. (<1%), which is a significant disadvantage for potential com. applications such as drug delivery and coatings. In principle, this problem can be circumvented by polymn.-induced block copolymer self-assembly. Here the authors detail the synthesis and subsequent in situ self-assembly of amphiphilic AB diblock copolymers in a one pot concd. aq. dispersion polymn. formulation. The authors show that spherical micelles, wormlike micelles, and vesicles can be predictably and efficiently obtained (within 2 h of polymn., >99% monomer conversion) at relatively high solids in purely aq. soln. Furthermore, careful monitoring of the in situ polymn. by transmission electron microscopy reveals various novel intermediate structures (including branched worms, partially coalesced worms, nascent bilayers, "octopi", "jellyfish", and finally pure vesicles) that provide important mechanistic insights regarding the evolution of the particle morphol. during the sphere-to-worm and worm-to-vesicle transitions. This environmentally benign approach (which involves no toxic solvents, is conducted at relatively high solids, and requires no addnl. processing) is readily amenable to industrial scale-up, since it is based on com. available starting materials.
- 19Lovett, J. R.; Ratcliffe, L. P. D.; Warren, N. J.; Armes, S. P.; Smallridge, M. J.; Cracknell, R. B.; Saunders, B. R. A Robust Cross-Linking Strategy for Block Copolymer Worms Prepared via Polymerization-Induced Self-Assembly. Macromolecules 2016, 49 (8), 2928– 2941, DOI: 10.1021/acs.macromol.6b00422Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XmtVWisL0%253D&md5=6c72e71bec9902b7aab48eb278fcdb53A Robust Cross-Linking Strategy for Block Copolymer Worms Prepared via Polymerization-Induced Self-AssemblyLovett, J. R.; Ratcliffe, L. P. D.; Warren, N. J.; Armes, S. P.; Smallridge, M. J.; Cracknell, R. B.; Saunders, B. R.Macromolecules (Washington, DC, United States) (2016), 49 (8), 2928-2941CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A poly(glycerol monomethacrylate) (PGMA) chain transfer agent is chain-extended by reversible addn.-fragmentation chain transfer (RAFT) statistical copolymn. of 2-hydroxypropyl methacrylate (HPMA) with glycidyl methacrylate (GlyMA) in concd. aq. soln. via polymn.-induced self-assembly (PISA). A series of five free-standing worm gels is prepd. by fixing the overall d.p. of the core-forming block at 144 while varying its GlyMA content from 0 to 20 mol %. 1H NMR kinetics indicated that GlyMA is consumed much faster than HPMA, producing a GlyMA-rich sequence close to the PGMA stabilizer block. Temp.-dependent oscillatory rheol. studies indicate that increasing the GlyMA content leads to progressively less thermoresponsive worm gels, with no degelation on cooling being obsd. for worms contg. 20 mol % GlyMA. The epoxy groups in the GlyMA residues can be ring-opened using 3-aminopropyltriethoxysilane (APTES) in order to prep. core crosslinked worms via hydrolysis-condensation with the siloxane groups and/or hydroxyl groups on the HPMA residues. Perhaps surprisingly, 1H NMR anal. indicates that the epoxy-amine reaction and the intermol. crosslinking occur on similar time scales. Crosslinking leads to stiffer worm gels that do not undergo degelation upon cooling. Dynamic light scattering studies and TEM analyses conducted on linear worms exposed to either methanol (a good solvent for both blocks) or anionic surfactant result in immediate worm dissocn. In contrast, crosslinked worms remain intact under such conditions, provided that the worm cores comprise at least 10 mol % GlyMA.
- 20Blanazs, A.; Ryan, A. J.; Armes, S. P. Predictive Phase Diagrams for RAFT Aqueous Dispersion Polymerization: Effect of Block Copolymer Composition, Molecular Weight, and Copolymer Concentration. Macromolecules 2012, 45 (12), 5099– 5107, DOI: 10.1021/ma301059rGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XosFemtrc%253D&md5=5847f249ec070800f649a2a2f13a2b86Predictive Phase Diagrams for RAFT Aqueous Dispersion Polymerization: Effect of Block Copolymer Composition, Molecular Weight, and Copolymer ConcentrationBlanazs, A.; Ryan, A. J.; Armes, S. P.Macromolecules (Washington, DC, United States) (2012), 45 (12), 5099-5107CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Polymn.-induced self-assembly (PISA) of poly(glycerol monomethacrylate)-poly(2-hydroxypropyl methacrylate) (PGMA-PHPMA) diblocks is conducted using a RAFT aq. dispersion polymn. formulation at 70 °C. Several PGMA macromol. chain transfer agents (macro-CTAs) are chain-extended using a water-miscible monomer (HPMA): the growing PHPMA block becomes increasingly hydrophobic and hence drives in situ self-assembly. The final copolymer morphol. in such PISA syntheses depends on just three parameters: the mean d.p. (DP) of the PGMA stabilizer block, the mean DP of the PHPMA core-forming block, and the total solids concn. TEM is used to construct detailed diblock copolymer phase diagrams for PGMA DPs of 47, 78, and 112. For the shortest stabilizer block, there is essentially no concn. dependence: spheres, worms, or vesicles can be obtained even at 10% wt./wt. solids simply by selecting the DP of the PHPMA block that gives the appropriate mol. curvature. For a PGMA DP of 78, the phase diagram is rich: and the copolymer morphol. depends strongly on the total solids concn. There is also a narrow region where spheres, worms, and vesicles coexist, which may be due to the effect of polydispersity. For a PGMA112 macro-CTA, the phase diagram is dominated by spherical morphologies. This is probably because the longer core-forming block DPs required to reduce the mol. curvature are significantly more dehydrated and hence less mobile, which prevents the in situ evolution of morphol. from spheres to higher order morphologies. This hypothesis is supported by the observation that addn. of ethanol to aq. PISA syntheses conducted using the longer macro-CTAs allows access to diblock copolymer worms or vesicles, since this cosolvent solvates the core-forming PHPMA chains and hence increases their mobility at 70 °C. Elucidation of such phase diagrams is vital to ensure reproducible targeting of pure phases, rather than mixed phases.
- 21Mable, C. J.; Gibson, R. R.; Prevost, S.; McKenzie, B. E.; Mykhaylyk, O. O.; Armes, S. P. Loading of Silica Nanoparticles in Block Copolymer Vesicles during Polymerization-Induced Self-Assembly: Encapsulation Efficiency and Thermally Triggered Release. J. Am. Chem. Soc. 2015, 137 (51), 16098– 16108, DOI: 10.1021/jacs.5b10415Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFSmtb%252FK&md5=3cc548f55bc0f03245e24924f8657210Loading of Silica Nanoparticles in Block Copolymer Vesicles during Polymerization-Induced Self-Assembly: Encapsulation Efficiency and Thermally Triggered ReleaseMable, Charlotte J.; Gibson, Rebecca R.; Prevost, Sylvain; McKenzie, Beulah E.; Mykhaylyk, Oleksandr O.; Armes, Steven P.Journal of the American Chemical Society (2015), 137 (51), 16098-16108CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Poly(glycerol monomethacrylate)-poly(2-hydroxypropyl methacrylate) diblock copolymer vesicles can be prepd. in the form of concd. aq. dispersions via polymn.-induced self-assembly (PISA). In the present study, these syntheses are conducted in the presence of varying amts. of silica nanoparticles of approx. 18 nm diam. This approach leads to encapsulation of up to hundreds of silica nanoparticles per vesicle. Silica has high electron contrast compared to the copolymer which facilitates TEM anal., and its thermal stability enables quantification of the loading efficiency via thermogravimetric anal. Encapsulation efficiencies can be calcd. using disk centrifuge photosedimentometry, since the vesicle d. increases at higher silica loadings while the mean vesicle diam. remains essentially unchanged. Small angle X-ray scattering (SAXS) is used to confirm silica encapsulation, since a structure factor is obsd. at q ≈ 0.25 nm-1. A new two-population model provides satisfactory data fits to the SAXS patterns and allows the mean silica vol. fraction within the vesicles to be detd. Finally, the thermoresponsive nature of the diblock copolymer vesicles enables thermally triggered release of the encapsulated silica nanoparticles simply by cooling to 0-10 °C, which induces a morphol. transition. These silica-loaded vesicles constitute a useful model system for understanding the encapsulation of globular proteins, enzymes, or antibodies for potential biomedical applications. They may also serve as an active payload for self-healing hydrogels or repair of biol. tissue. Finally, we also encapsulate a model globular protein, bovine serum albumin, and calc. its loading efficiency using fluorescence spectroscopy.
- 22Rösler, A.; Vandermeulen, G. W. M.; Klok, H. A. Advanced Drug Delivery Devices via Self-Assembly of Amphiphilic Block Copolymers. Adv. Drug Delivery Rev. 2012, 64, 270– 279, DOI: 10.1016/j.addr.2012.09.026Google ScholarThere is no corresponding record for this reference.
- 23Derry, M. J.; Mykhaylyk, O. O.; Armes, S. P. A Vesicle-to-Worm Transition Provides a New High-Temperature Oil Thickening Mechanism. Angew. Chemie - Int. Ed. 2017, 56 (7), 1746– 1750, DOI: 10.1002/anie.201609365Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlsFCjtg%253D%253D&md5=c21c3e5268d7378a9d75321f947e0589A Vesicle-to-Worm Transition Provides a New High-Temperature Oil Thickening MechanismDerry, Matthew J.; Mykhaylyk, Oleksandr O.; Armes, Steven P.Angewandte Chemie, International Edition (2017), 56 (7), 1746-1750CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Diblock copolymer vesicles are prepd. via RAFT dispersion polymn. directly in mineral oil. Such vesicles undergo a vesicle-to-worm transition on heating to 150 °C, as judged by TEM and SAXS. Variable-temp. 1H NMR spectroscopy indicates that this transition is the result of surface plasticization of the membrane-forming block by hot solvent, effectively increasing the vol. fraction of the stabilizer block and so reducing the packing parameter for the copolymer chains. The rheol. behavior of a 10 % wt./wt. copolymer dispersion in mineral oil is strongly temp.-dependent: the storage modulus increases by five orders of magnitude on heating above the crit. gelation temp. of 135 °C, as the non-interacting vesicles are converted into weakly interacting worms. SAXS studies indicate that, on av., three worms are formed per vesicle. Such vesicle-to-worm transitions offer an interesting new mechanism for the high-temp. thickening of oils.
- 24Bates, F. S. Polymer-Polymer Phase Behavior. Science (80-.). 1991, 251 (4996), 898– 905, DOI: 10.1126/science.251.4996.898Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhsVCgsrk%253D&md5=53fc355f97cf74f3bf5e17e2c427509fPolymer-polymer phase behaviorBates, Frank S.Science (Washington, DC, United States) (1991), 251 (4996), 898-905CODEN: SCIEAS; ISSN:0036-8075.A review with 57 refs. on basic factors governing polymer-polymer phase behavior with emphasis on linear homopolymer blends and diblock copolymers as representative model mol. architectures. Equil. thermodn. and phase sepn. dynamics are reviewed and discussed.
- 25Blanazs, A.; Armes, S. P.; Ryan, A. J. Self-Assembled Block Copolymer Aggregates: From Micelles to Vesicles and Their Biological Applications. Macromol. Rapid Commun. 2009, 30 (4), 267– 277, DOI: 10.1002/marc.200800713Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXivFelu7k%253D&md5=a7578981e5867fcee9d50b9e361c9597Self-assembled block copolymer aggregates: from micelles to vesicles and their biological applicationsBlanazs, Adam; Armes, Steven P.; Ryan, Anthony J.Macromolecular Rapid Communications (2009), 30 (4-5), 267-277CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The ability of amphiphilic block copolymers to self-assemble in selective solvents has been widely studied in academia and utilized for various com. products. The self-assembled polymer vesicle is at the forefront of this nanotechnol. revolution with seemingly endless possible uses, ranging from biomedical to nanometer-scale enzymic reactors. This review is focused on the inherent advantages in using polymer vesicles over their small mol. lipid counterparts and the potential applications in biol. for both drug delivery and synthetic cellular reactors.
- 26Albigès, R.; Klein, P.; Roi, S.; Stoffelbach, F.; Creton, C.; Bouteiller, L.; Rieger, J. Water-Based Acrylic Coatings Reinforced by PISA-Derived Fibers. Polym. Chem. 2017, 8 (34), 4992– 4995, DOI: 10.1039/C7PY00302AGoogle Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmtFGnt70%253D&md5=7c7acccbb5661d2aad39d9b2b74729d3Water-based acrylic coatings reinforced by PISA-derived fibersAlbiges, Richard; Klein, Pauline; Roi, Stephanie; Stoffelbach, Francois; Creton, Costantino; Bouteiller, Laurent; Rieger, JuttaPolymer Chemistry (2017), 8 (34), 4992-4995CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Polymn.-induced self-assembly (PISA) is an efficient approach to synthesize self-assembled core-shell nanofibers in water, made of amphiphilic block copolymers. We demonstrate here for the first time that such anisotropic and high Tg nanofibers can be used as reinforcing fillers for water-based acrylic materials. This strategy opens the way for strong all-org. aq. coatings.
- 27Chenal, M.; Rieger, J.; Vechambre, C.; Chenal, J.-M.; Chazeau, L.; Creton, C.; Bouteiller, L. Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard Phase. Macromol. Rapid Commun. 2013, 34 (34), 1524– 1529, DOI: 10.1002/marc.201300410Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlahtLbO&md5=4d4bc64299888278198f8cdf39b17d57Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard PhaseChenal, Marion; Rieger, Jutta; Vechambre, Cyril; Chenal, Jean-Marc; Chazeau, Laurent; Creton, Costantino; Bouteiller, LaurentMacromolecular Rapid Communications (2013), 34 (19), 1524-1529CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)In this study, aq. emulsion polymn. of Bu acrylate is performed in batch conditions without surfactants using a poly(acrylic acid)-trithiocarbonate macro-RAFT agent to control the polymn. and to stabilize the emulsion. According to the polymn.-induced self-assembly (PISA) approach, well-defined amphiphilic PAA-b-PBA diblock copolymers form and self-assemble during synthesis to yield highly stable core-shell particles with an extremely thin hard PAA shell. For the first time, we report here the specific properties of films obtained from these particular latexes. After drying the aq. dispersion, tough and transparent films are obtained. Although the films are not chem. crosslinked, they do not dissolve in good solvents for PBA. Moreover, they remain transparent even after immersion in water. Rheol. shows that the films are both stiff and ductile, thanks to the nanostructured but very low vol. fraction (less than 3 wt%) of PAA forming a percolating network in the soft PBA. Compared with conventional core-shell-based films, this approach affords for the first time a route to a thin percolating honeycomb nanostructure with a sharp and strong interface between the two phases. The versatility of the synthetic procedure opens perspectives for a large range of functional materials.
- 28Rieger, J.; Osterwinter, G.; Bui, C.; Stoffelbach, F.; Charleux, B. Surfactant-Free Controlled/Living Radical Emulsion (Co)Polymerization of n-Butyl Acrylate and Methyl Methacrylate via RAFT Using Amphiphilic Polyethylene Oxide-Based Trithiocarbonate Chain Transfer Agents. Macromolecules 2009, 42 (15), 5518– 5525, DOI: 10.1021/ma9008803Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXotVGlt74%253D&md5=d13143067792442f9941957db3ccc8d9Surfactant-Free Controlled/Living Radical Emulsion (Co)polymerization of n-Butyl Acrylate and Methyl Methacrylate via RAFT Using Amphiphilic Poly(ethylene oxide)-Based Trithiocarbonate Chain Transfer AgentsRieger, Jutta; Osterwinter, Gregor; Bui, Chuong; Stoffelbach, Francois; Charleux, BernadetteMacromolecules (Washington, DC, United States) (2009), 42 (15), 5518-5525CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The RAFT-mediated, surfactant-free, ab initio, batch emulsion polymn. of Bu acrylate (nBA) and its copolymn. with Me methacrylate (MMA) were studied. The control agent was a surface-active trithiocarbonate macromol. RAFT agent composed of a hydrophilic poly(ethylene oxide) (PEO) block and a hydrophobic dodecyl chain. The homopolymns. of nBA were fast with high final conversions, and the polymer chains were well-controlled with narrow molar mass distribution. The length of the PEO chain was shown to affect the particle size and the polymn. kinetics directly. We found the conditions to tune the particle size independently from the poly(Bu acrylate) chain length by playing with a mixt. of macro-RAFT agents with long and short PEO segment or by adding a PEO-based nonionic surfactant. The copolymns. of nBA and MMA exhibited features very similar to those of the nBA homopolymns. provided that the molar percentage of MMA did not exceed approx. 75%. In all cases, stable, submicrometric particles composed of amphiphilic diblock copolymer chains were formed.
- 29Cockram, A. A.; Bradley, R. D.; Lynch, S. A.; Fleming, P. C. D.; Williams, N. S. J.; Murray, M. W.; Emmett, S. N.; Armes, S. P. Optimization of the High-Throughput Synthesis of Multiblock Copolymer Nanoparticles in Aqueous Media: Via Polymerization-Induced Self-Assembly. React. Chem. Eng. 2018, 3 (5), 645– 657, DOI: 10.1039/C8RE00066BGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtFCjsrfF&md5=b4951aaeaec274ad421b06f6c01670afOptimization of the high-throughput synthesis of multiblock copolymer nanoparticles in aqueous media via polymerization-induced self-assemblyCockram, Amy A.; Bradley, Robert D.; Lynch, Sylvie A.; Fleming, Patricia C. D.; Williams, Neal S. J.; Murray, Martin W.; Emmett, Simon N.; Armes, Steven P.Reaction Chemistry & Engineering (2018), 3 (5), 645-657CODEN: RCEEBW; ISSN:2058-9883. (Royal Society of Chemistry)In the present study, we report that PISA formulations are sufficiently robust to enable high-throughput expts. using a com. synthesis robot (Chemspeed Autoplant A100). More specifically, we use reversible addn.-fragmentation chain transfer (RAFT) aq. emulsion polymn. of either Bu methacrylate and/or benzyl methacrylate to prep. various examples of methacrylic multiblock copolymer nanoparticles using a poly(methacrylic acid) stabilizer block. Adequate stirring is essential to generate sufficiently small monomer droplets for such heterogeneous polymns. to proceed efficiently. Good reproducibility can be achieved under such conditions, with well-defined spherical morphologies being obtained at up to 45% wt./wt. solids. GPC studies indicate high blocking efficiencies but relatively broad mol. wt. distributions (Mw/Mn = 1.36-1.85), suggesting well-defined (albeit rather polydisperse) block copolymer chains. These preliminary studies provide a sound basis for high-throughput screening of RAFT-mediated PISA formulations, which is likely to be required for commercialization of this technol. Our results indicate that methacrylic PISA formulations enable the synthesis of diblock and triblock copolymer nanoparticles in high overall yield (94-99%) within 1-3 h at 70 °C. However, tetrablocks suffer from incomplete conversions (87-96% within 5 h) and hence most likely represent the upper limit for this approach.
- 30Cockram, A. A.; Neal, T. J.; Derry, M. J.; Mykhaylyk, O. O.; Williams, N. S. J.; Murray, M. W.; Emmett, S. N.; Armes, S. P. Effect of Monomer Solubility on the Evolution of Copolymer Morphology during Polymerization-Induced Self-Assembly in Aqueous Solution. Macromolecules 2017, 50 (3), 796– 802, DOI: 10.1021/acs.macromol.6b02309Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsFKms7w%253D&md5=cd3a26b52453e4fa9ab4f727e5b8a9feEffect of Monomer Solubility on the Evolution of Copolymer Morphology during Polymerization-Induced Self-Assembly in Aqueous SolutionCockram, Amy A.; Neal, Thomas J.; Derry, Matthew J.; Mykhaylyk, Oleksandr O.; Williams, Neal S. J.; Murray, Martin W.; Emmett, Simon N.; Armes, Steven P.Macromolecules (Washington, DC, United States) (2017), 50 (3), 796-802CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Polymn.-induced self-assembly (PISA) has become a widely used technique for the design of diblock copolymer nano-objects in concd. aq. soln. Depending on the specific PISA formulation, reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. typically provides straightforward access to either spheres, worms or vesicles. In contrast, RAFT aq. emulsion polymn. formulations often lead to just kinetically trapped spheres. This limitation is currently not understood and only a few empirical exceptions were reported in the literature. In the present work, the effect of monomer soly. on copolymer morphol. is explored for an aq. polymn.-induced self-assembly (PISA) formulation. Using 2-hydroxybutyl methacrylate (aq. soly. =20 g dm-3 at 70 °C) instead of benzyl methacrylate (0.40 g dm-3 at 70 °C) for the core-forming block allows access to an unusual "monkey nut" copolymer morphol. over a relatively narrow range of target ds.p. when using a poly(methacrylic acid) RAFT agent at pH 5. These new nanoparticles have been characterized by transmission electron microscopy, dynamic light scattering, aq. electrophoresis, shear-induced polarized light imaging (SIPLI) and small-angle X-ray scattering.
- 31Rieger, J. Guidelines for the Synthesis of Block Copolymer Particles of Various Morphologies by RAFT Dispersion Polymerization. Macromol. Rapid Commun. 2015, 36 (16), 1458– 1471, DOI: 10.1002/marc.201500028Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXovVCku78%253D&md5=bef40a4da086930c94c52099f796b0bcGuidelines for the Synthesis of Block Copolymer Particles of Various Morphologies by RAFT Dispersion PolymerizationRieger, JuttaMacromolecular Rapid Communications (2015), 36 (16), 1458-1471CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This article presents the recent developments of radical dispersion polymerizaton controlled by reversible addn. fragmentation chain transfer (RAFT) for the prodn. of block copoly-mer particles of various morphologies, such as core-shell spheres, worms, or vesicles. It is not meant to be an exhaustive review but it rather provides guidelines for non-specialists. The article is subdivided into eight sections. After a general introduction, the mechanism of polymn.-induced self-assembly (PISA) through RAFT-mediated dispersion polymn. is presented and the different parameters that control the morphol. produced are discussed. The next two sections are devoted to the choice of the monomer/solvent pair and the macroRAFT agent. Afterwards, post-polymn. morphol. order-to-order transitions (i.e. morphol. transitions triggered by extrinsic stimuli) or order-to-disorder transitions (i.e. disassembly of chains) are discussed. Assemblies based on more complex polymer architectures, such as triblock copolymers, are presented next, and finally the possibility to stabilize these structures by crosslinking is reported. The manuscript ends with a short conclusion and an outlook.
- 32Penfold, N. J. W.; Whatley, J. R.; Armes, S. P. Thermoreversible Block Copolymer Worm Gels Using Binary Mixtures of PEG Stabilizer Blocks. Macromolecules 2019, 52 (4), 1653– 1662, DOI: 10.1021/acs.macromol.8b02491Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXis1antbw%253D&md5=0f5272986cdd505e3a655d03c7d37140Thermoreversible Block Copolymer Worm Gels Using Binary Mixtures of PEG Stabilizer BlocksPenfold, Nicholas J. W.; Whatley, Jessica R.; Armes, Steven P.Macromolecules (Washington, DC, United States) (2019), 52 (4), 1653-1662CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Two trithiocarbonate-based poly(ethylene glycol) (PEG) macromol. chain transfer agents (macro-CTAs) with mean ds.p. of 45 and 113 were prepd. with ≥94% chain-end functionality. Binary mixts. of these PEG-trithiocarbonate macro-CTAs were then chain-extended via reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. of 2-hydroxypropyl methacrylate (HPMA). Systematic variation of the relative proportions of PEG45 and PEG113 macro-CTAs and the d.p. of the PHPMA core-forming block resulted in the formation of [x PEG45 + z PEG113] - PHPMAn block copolymer spheres, worms, or vesicles, where x and z represent the mole fractions of PEG45 and PEG113, resp. A phase diagram was constructed to establish the relationship between block copolymer compn. and nanoparticle morphol. The thermoresponsive behavior of block copolymer worms was assessed by visual inspection, dynamic light scattering (DLS), transmission electron microscopy (TEM) and temp.-dependent oscillatory rheol. Increasing the proportion of PEG45 (x = 0.00-0.40) in the stabilizer block resulted in a moderate increase in worm gel strength, but cooling resulted in irreversible degelation owing to a worm-to-sphere morphol. transition. However, the phase diagram enabled identification of a single diblock copolymer compn. that exhibited reversible degelation behavior in pure water. This formulation was then further optimized to exhibit the same rheol. behavior in a com. cell culture medium (Nutristem) by fixing the PEG mole fraction at x = 0.70 while lowering the PHPMA DP from 115 to 75. Importantly, the gel strength at physiol. temp. can be readily tuned simply by variation of the copolymer concn. In principle, this study has important implications for the preservation of human stem cells, which can enter stasis when immersed in certain worm gels [see: Canton et al. ACS Cent. Sci.2016, 2, 65-74].
- 33Qiao, Z.; Qiu, T.; Liu, W.; Zhang, L.; Tu, J.; Guo, L.; Li, X. A “Green” Method for Preparing ABCBA Penta-Block Elastomers by Using RAFT Emulsion Polymerization. Polym. Chem. 2017, 8 (19), 3013– 3021, DOI: 10.1039/C7PY00464HGoogle Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmtVSis7w%253D&md5=2bfb612e9ce191da9a20f7d2843f94feA "green" method for preparing ABCBA penta-block elastomers by using RAFT emulsion polymerizationQiao, Zhi; Qiu, Teng; Liu, Weiwei; Zhang, Liangdong; Tu, Jinqiang; Guo, Longhai; Li, XiaoyuPolymer Chemistry (2017), 8 (19), 3013-3021CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A clean method for prepg. ABCBA penta-block terpolymers was developed in a surfactant and org. solvent free emulsion system via one-pot RAFT polymn. The prepd. ABCBA penta-block terpolymers show large fracture tensile strain, low permanent set and potential application as elastomer materials. During the prepn. process, polyacrylic acid (PAA) was firstly synthesized as the A block via RAFT polymn. in water in the presence of a dual-functional RAFT agent. Subsequently, a partially water-sol. monomer 2,2,2-trifluoroethyl acrylate (TFEA) was used to synthesize the B block, and ABA tri-block copolymers could be obtained via org. solvent free RAFT emulsion polymn. Using the ABA tri-block copolymers as a surfactant, the ABCBA penta-block terpolymers were further synthesized using 2,2,3,4,4,4-hexafluorobutyl acrylate (HFBA) as the third monomer in an org. solvent free system. It has been found that the emulsification efficiency of the ABA tri-block copolymers is higher than that of the corresponding AB di-block copolymers. The assembly mechanism of penta-block terpolymers to form a tri-layer core-shell latex particle was proposed and confirmed by TEM images. Finally, a transparent and hydrophobic elastomeric film was obtained by directly casting the synthesized penta-block terpolymer latexes and subsequent heat treatment. The films have excellent mech. properties and water resistance.
- 34Wu, X.; Qiao, Y.; Yang, H.; Wang, J. Self-Assembly of a Series of Random Copolymers Bearing Amphiphilic Side Chains. J. Colloid Interface Sci. 2010, 349 (2), 560– 564, DOI: 10.1016/j.jcis.2010.05.093Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXpt1Sktrw%253D&md5=f075923332cfb1412f93b0aab6d75fd5Self-assembly of a series of random copolymers bearing amphiphilic side chainsWu, Xu; Qiao, Yingjie; Yang, Hui; Wang, JinbenJournal of Colloid and Interface Science (2010), 349 (2), 560-564CODEN: JCISA5; ISSN:0021-9797. (Elsevier B.V.)A novel series of comb-like random copolymers were prepd. by polymn. of amphiphilic macromonomers, 2-(acrylamido)-octane sulfonic acid (AMC8S), 2-(acrylamido)-dodecane sulfonic acid (AMC12S), and 2-(acrylamido)-hexadecane sulfonic acid (AMC16S), with 2-(acrylamido)-2-methylpropanesulfonic acid (AMPS) resp. The synthesis of the polymers with the same contents of amphiphilic units as side chains, but different chain length, enabled us to study the chain length dependence of their assocn. in salt soln. Steady-state fluorescence measurements with pyrene as a polarity probe, quasielastic light scattering techniques (QELS) and transmission electron micrograph (TEM) were employed to investigate the associative properties of the system. The above investigations showed that all kinds of side chains begin to assemble at certain polymer concns. and the crit. aggregation concn. (CAC) decrease dramatically with the increase in the length and content of alkyl. An interesting phenomenon is that the assembly tends more favorably to occur among different mols. rather than within single mol. when the no. of carbon atoms in the alkyl groups or the polymer concn. increases, leading to the formation of larger multimol. micelle-like aggregate. The aim of the present work is to establish the fundamental preconditions of intramol. and intermol. assocn. fashions for the polymers, which is useful for the exploitation of functional groups and contributes to the development of amphiphilic random polymers.
- 35Kawata, T.; Hashidzume, A.; Sato, T. Micellar Structure of Amphiphilic Statistical Copolymers Bearing Dodecyl Hydrophobes in Aqueous Media. Macromolecules 2007, 40 (4), 1174– 1180, DOI: 10.1021/ma062299xGoogle Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXmtFSktg%253D%253D&md5=96d988e993768de20f3d0805aaedcc4cMicellar Structure of Amphiphilic Statistical Copolymers Bearing Dodecyl Hydrophobes in Aqueous MediaKawata, Takefumi; Hashidzume, Akihito; Sato, TakahiroMacromolecules (2007), 40 (4), 1174-1180CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The structure of micellar aggregates formed from ionic statistical copolymers of N-acryloyl-amino acids and N-dodecylmethacrylamide in 0.05 M aq. NaCl was studied by light scattering and fluorescence. The exptl. results indicated that the tendency of the interchain aggregation increased with the hydrophobic monomer content for each series of the copolymers but decreased with hydrophobicity of the amino acid residue in the copolymers. On the other hand, while the micellar structure of the statistical copolymers strongly depended on the kind of the hydrophobe, it is little dependent on the kind of the electrolyte monomer unit at low ionic strength. Using these data, a theor. anal. taking into account the chain stiffness effect revealed that uni-core micelles formed from the ionic statistical copolymers were of flower type with a min. loop size.
- 36Zhu, X.; Liu, M. Self-Assembly and Morphology Control of New l -Glutamic Acid-Based Amphiphilic Random Copolymers: Giant Vesicles, Vesicles, Spheres, and Honeycomb Film. Langmuir 2011, 27 (21), 12844– 12850, DOI: 10.1021/la202680jGoogle Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXht1OrtrbN&md5=84bfb08f30edb16a065a4d5e0292d0a3Self-Assembly and Morphology Control of New L-Glutamic Acid-Based Amphiphilic Random Copolymers: Giant Vesicles, Vesicles, Spheres, and Honeycomb FilmZhu, Xuewang; Liu, MinghuaLangmuir (2011), 27 (21), 12844-12850CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)New amphiphilic random copolymers contg. hydrophobic dodecyl (C12) chain and hydrophilic L-glutamic acid were synthesized, and their self-assembly in soln. as well as on the solid surfaces was investigated. The self-assembly behavior of these polymers are largely dependent on their hydrophilic and hydrophobic balances. The copolymer with a more hydrophobic alkyl chain (∼90%) self-assembled into giant vesicles with a diam. of several micrometers in a mixed solvent of ethanol and water. When the hydrophobic ratio decreased to ca. 76%, the polymer self-assembled into conventional vesicles with several hundred nanometers. The giant vesicles could be fused in certain conditions, while the conventional vesicles were stable. When the content of the hydrophilic part was further increased, no organized structures were formed. On the other hand, when the copolymer solns. were directly cast on solid substrates such as silicon plates, films with organized nanostructures could also be obtained, the morphol. of which depended on solvent selection. When ethanol or methanol was used, spheres were obtained. When dichloromethane was used as the solvent, honeycomb-like morphologies were obtained. These results showed that through appropriate mol. design, random copolymer could self-assemble into various organized structures, which could be regulated through the hydrophobic/hydrophilic balance and the solvents.
- 37Tian, F.; Yu, Y.; Wang, C.; Yang, S. Consecutive Morphological Transitions in Nanoaggregates Assembled from Amphiphilic Random Copolymer via Water Driven Micellization and Light-Triggered Dissociation. Macromolecules 2008, 41 (10), 3385– 3388, DOI: 10.1021/ma800142jGoogle Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltlalsro%253D&md5=1c16592aaf10bede6627e031b226fe94Consecutive Morphological Transitions in Nanoaggregates Assembled from Amphiphilic Random Copolymer via Water-Driven Micellization and Light-Triggered DissociationTian, Feng; Yu, Yuanyuan; Wang, Changchun; Yang, ShuMacromolecules (Washington, DC, United States) (2008), 41 (10), 3385-3388CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)We report a study of continuous morphol. transitions from spherical micelles, through hollow tubes and wormlike rods, to large vesicles in nanoaggregates self-assembled from photoresponsive amphiphilic random copolymer. Through esterification of the side chain hydroxyl groups, the hydrophilic poly(hydroxyethyl methacrylate) (PHEMA) backbone is partially and randomly modified by the hydrophobic, light-responsive 2-diazo-1,2-naphthoquinone (DNQ) mols. The micellization and transition are then triggered simply by adding water into the copolymer/DMF soln. Upon irradn. to UV light at 405 nm, large vesicles become increasingly hydrophilic with time and gradually dissoc. into globular hydrogel particles in the aq. soln. (Scheme 1), releasing the encapsulated dye mols.
- 38Sun, G.; Zhang, M.; He, J.; Ni, P. Synthesis of Amphiphilic Cationic Copolymers Poly[2-(Methacryloyloxy)Ethyl Trimethylammonium Chloride-co-stearyl Methacrylate] and Their Self-Assembly Behavior in Water and Water-Ethanol Mixtures. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 4670– 4684, DOI: 10.1002/pola.23517Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXpvFCksb0%253D&md5=72e4ba806b4ccb2973547000d316ec1bSynthesis of amphiphilic cationic copolymers poly[2-(methacryloyloxy)ethyl trimethylammonium chloride-co-stearyl methacrylate] and their self-assembly behavior in water and water-ethanol mixturesSun, Guixiang; Zhang, Mingzu; He, Jinlin; Ni, PeihongJournal of Polymer Science, Part A: Polymer Chemistry (2009), 47 (18), 4670-4684CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A series of amphiphilic cationic random copolymers, namely poly[2-(methacryloyloxy)ethyl trimethylammonium chloride-co-stearyl methacrylate] or poly(MADQUAT-co-SMA), were synthesized via conventional free-radical copolymn. using 2,2'-azobisisobutyronitrile (AIBN) as initiator and n-dodecanethiol as chain transfer agent. The resultant products were then characterized by FT-IR, 1H NMR, MALDI-TOF MS measurements. From the no.-av. mol. wts. of the copolymers, we can conclude that these copolymers have oligomeric structure with a limited no. of hydrophilic and hydrophobic moieties in a short polymer chain. The reactivity ratios (rMADQUAT = 0.83, rSMA = 0.25) between the hydrophilic MADQUAT monomer and the hydrophobic SMA monomer were calcd. by the Finemann and Ross method, which was based on the results of 1H NMR anal. The surface activity of the random copolymers was studied by the combination of surface tension and contact angle measurement, and these copolymers possess relatively high surface activity. The crit. aggregation concns. (cac) of the copolymers in aq. soln. were detd. by fluorescence probe method and surface tension measurement. The different nanoparticles of poly(MADQUAT-co-SMA) copolymers formed in pure water or ethanol-water mixt. were proved by the particle size and size distribution in the measurement of dynamic light scattering (DLS). Furthermore, using TEM, we could observe various self-assembly morphologies of these random copolymer. All the amphiphilic cationic random copolymers have a good self-assembly behavior, even if they are ill-defined copolymers.
- 39Stephan, T.; Muth, S.; Schmidt, M. Shape Changes of Statistical Copolymacromonomers: From Wormlike Cylinders to Horseshoe- and Meanderlike Structures. Macromolecules 2002, 35 (27), 9857– 9860, DOI: 10.1021/ma025711rGoogle Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XptFOks74%253D&md5=df799d57f47cf53452c936948493ef19Shape Changes of Statistical Copolymacromonomers: From Wormlike Cylinders to Horseshoe- and Meander-like StructuresStephan, Tim; Muth, Sandra; Schmidt, ManfredMacromolecules (2002), 35 (27), 9857-9860CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Depending on the solvent, statistical copolymer cylindrical brushes adopt different shapes from worm-like to horseshoe or meander-like structures when spin-cast onto mica. To the best of our knowledge, this is the first time that shape persistent synthetic macromols. could be chem. manipulated to change their conformation from a wormlike to regularly curved structures. It should be mentioned, though, that for cylindrical brushes with Bu acrylate side chains a collapse of the wormlike conformation to spherical structures could be induced under pressure on a Langmuir-Blodgett trough. Also, poly(p-phenylene)s with amphipolar dendritic side chains were reported to have the potential to intramol. segregate lengthwise, particularly when spread on a LB trough as well. Our next step will address the question of whether the collapse of the minority component could be detected by scattering expts. in soln., which, however, was not successful so far.
- 40Ilhan, F.; Galow, T. H.; Gray, M.; Clavier, G.; Rotello, V. M. Giant Vesicle Formation through Self-Assembly of Complementary Random Copolymers. J. Am. Chem. Soc. 2000, 122 (24), 5895– 5896, DOI: 10.1021/ja0011966Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjs1elt7Y%253D&md5=86ebd81f301c14d25ad679cf7e6098e5Giant Vesicle Formation through Self-Assembly of Complementary Random CopolymersIlhan, Faysal; Galow, Trent H.; Gray, Mark; Clavier, Gilles; Rotello, Vincent M.Journal of the American Chemical Society (2000), 122 (24), 5895-5896CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Differential interference contrast optical microscopy and confocal laser scanning microscopy showed that vesicular H-bond complexes formed between chains of 4-(chloromethyl)styrene-styrene copolymer (I)-2,6-bis(propionamido)pyrid-4-one adducts and I-thymine-1-acetic acid adducts in CHCl3 solns.
- 41Liu, X.; Kim, J.-S.; Wu, J.; Eisenberg, A. Bowl-Shaped Aggregates from the Self-Assembly of an Amphiphilic Random Copolymer of Poly(Styrene- Co -Methacrylic Acid). Macromolecules 2005, 38 (16), 6749– 6751, DOI: 10.1021/ma050665rGoogle Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmt1Krtrw%253D&md5=3d6e4ea1e6486b954e76bbc45a4b6fa2Bowl-Shaped Aggregates from the Self-Assembly of an Amphiphilic Random Copolymer of Poly(styrene-co-methacrylic acid)Liu, Xiaoya; Kim, Joon-Seop; Wu, Jun; Eisenberg, AdiMacromolecules (2005), 38 (16), 6749-6751CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A bowl-shaped structure of random poly(styrene-co-methacrylic acid) was prepd. by self-assembly in dil. soln. We emphasize that the copolymer was a random copolymer in contrast to the block copolymers or functionally terminated homopolymers used before. The added viscosity control is probably provided by H bond interactions among OH groups along the backbone as in encountered in carboxylic acids in bulk or in org. solvents.
- 42Neal, T. J.; Parnell, A. J.; King, S. M.; Beattie, D. L.; Murray, M. W.; Williams, N. S. J.; Emmett, S. N.; Armes, S. P.; Spain, S. G.; Mykhaylyk, O. O. Control of Particle Size in the Self-Assembly of Amphiphilic Statistical Copolymers. Macromolecules 2021, 54 (3), 1425– 1440, DOI: 10.1021/acs.macromol.0c02341Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhslehs7o%253D&md5=a59483ff3e5c24287beb7c07816f64dcControl of Particle Size in the Self-Assembly of Amphiphilic Statistical CopolymersNeal, Thomas J.; Parnell, Andrew J.; King, Stephen M.; Beattie, Deborah L.; Murray, Martin W.; Williams, Neal S. J.; Emmett, Simon N.; Armes, Steven P.; Spain, Sebastian G.; Mykhaylyk, Oleksandr O.Macromolecules (Washington, DC, United States) (2021), 54 (3), 1425-1440CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A range of amphiphilic statistical copolymers is synthesized where the hydrophilic component is either methacrylic acid (MAA) or 2-(dimethylamino)ethyl methacrylate (DMAEMA) and the hydrophobic component comprises Me, Et, Bu, hexyl, or 2-ethylhexyl methacrylate, which provide a broad range of partition coeffs. (log P). Small-angle x-ray scattering studies confirm that these amphiphilic copolymers self-assemble to form well-defined spherical nanoparticles in an aq. soln., with more hydrophobic copolymers forming larger nanoparticles. Varying the nature of the alkyl substituent also influenced self-assembly with more hydrophobic comonomers producing larger nanoparticles at a given copolymer compn. A model based on particle surface charge d. (PSC model) is used to describe the relation between copolymer compn. and nanoparticle size. This model assumes that the hydrophilic monomer is preferentially located at the particle surface and provides a good fit to all of the exptl. data. More specifically, a linear relation is obsd. between the surface area fraction covered by the hydrophilic comonomer required to achieve stabilization and the log P value for the hydrophobic comonomer. Contrast variation small-angle neutron scattering is used to study the internal structure of these nanoparticles. This technique indicates partial phase sepn. within the nanoparticles, with about half of the available hydrophilic comonomer repeat units being located at the surface and hydrophobic comonomer-rich cores. This information enables a refined PSC model to be developed, which indicates the same relation between the surface area fraction of the hydrophilic comonomer and the log P of the hydrophobic comonomer repeat units for the anionic (MAA) and cationic (DMAEMA) comonomer systems. This study demonstrates how nanoparticle size can be readily controlled and predicted using relatively ill-defined statistical copolymers, making such systems a viable attractive alternative to diblock copolymer nanoparticles for a range of industrial applications.
- 43Azuma, Y.; Terashima, T.; Sawamoto, M. Self-Folding Polymer Iron Catalysts for Living Radical Polymerization. ACS Macro Lett. 2017, 6 (8), 830– 835, DOI: 10.1021/acsmacrolett.7b00498Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhtF2hurnN&md5=57e0d5b1b3ad8706735d4be8edc6faf7Self-Folding Polymer Iron Catalysts for Living Radical PolymerizationAzuma, Yusuke; Terashima, Takaya; Sawamoto, MitsuoACS Macro Letters (2017), 6 (8), 830-835CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)Iron-bearing self-folding polymers were created with amphiphilic random copolymers as active, versatile, and recyclable polymer-supported catalysts for living radical polymn. (LRP). The key is to build bis(imino)pyridine ligand cavities for iron complexes as linking units within self-folding polymers. Self-folding polymer ligands are synthesized by the intramol. imine crosslinking of self-folded amphiphilic random copolymers bearing hydrophilic poly(ethylene glycol) (PEG), hydrophobic dodecyl, and urea/aniline pendants with 2,6-pyridinedicarboxaldehyde in water. The folding polymers efficiently formed iron complex catalysts in the cores to induce LRP and random or block copolymn. of various methacrylates. The self-folding polymer catalysts not only showed high activity and tolerance to functional groups such as acid, hydroxyl groups, and oxygen but also afforded easy product recovery and catalyst recycle thanks to hydrophilic PEG chains.
- 44Lyon, C. K.; Prasher, A.; Hanlon, A. M.; Tuten, B. T.; Tooley, C. A.; Frank, P. G.; Berda, E. B. A Brief User’s Guide to Single-Chain Nanoparticles. Polym. Chem. 2015, 6 (2), 181– 197, DOI: 10.1039/C4PY01217HGoogle Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVKmtbrP&md5=6018dc5ae1b7d54385638a6da6e6f4c2A brief user's guide to single-chain nanoparticlesLyon, Christopher K.; Prasher, Alka; Hanlon, Ashley M.; Tuten, Bryan T.; Tooley, Christian A.; Frank, Peter G.; Berda, Erik B.Polymer Chemistry (2015), 6 (2), 181-197CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)In this review we outline the various methods that have been explored to synthesize architecturally defined nanoparticles from discrete polymer chains, summarize the methods of characterization that are required to prove their formation and probe their morphol., and introduce a no. of potential applications that are being explored currently. Given the small size of the nanostructures produced by these methods and the relative ease with which they can be tailored to specific end use applications it is likely such efforts will intensify in the coming years. So far, simple chem. has been utilized and high-level characterization and modeling studies have been applied to understand the process by which these particles form and how they behave, both in the bulk and in soln. Although impossible to predict where this work will lead, we hope this "user's guide" will prove useful to the community as research on single-chain nanoparticles continues to evolve.
- 45Romulus, J.; Weck, M. Single-Chain Polymer Self-Assembly Using Complementary Hydrogen Bonding Units. Macromol. Rapid Commun. 2013, 34 (19), 1518– 1523, DOI: 10.1002/marc.201300501Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlSiu7bP&md5=f2a6a6fa1b58d398fc781d53ecebab8dSingle-Chain Polymer Self-Assembly Using Complementary Hydrogen Bonding UnitsRomulus, Joy; Weck, MarcusMacromolecular Rapid Communications (2013), 34 (19), 1518-1523CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A triblock copolymer contg. the complementary hydrogen bonding recognition pair ureidoguanosine-diaminonaphthyridine (UG-DAN) as pendant functional groups is synthesized using ring-opening metathesis polymn. (ROMP). The norbornene-based DAN monomer is shown to allow for a controlled polymn. when polymd. in the presence of a modified-UG mol. that serves as a protecting group, subsequently allowing for the fabrication of functionalized triblock copolymers. The self-assembly of the copolymers was characterized using dynamic light scattering and 1H NMR spectroscopy. It is demonstrated that the polymers self-assemble via complementary hydrogen bonding motifs even at low dilns., indicating intramol. interactions.
- 46Gillissen, M. A. J.; Voets, I. K.; Meijer, E. W.; Palmans, A. R. A. Single Chain Polymeric Nanoparticles as Compartmentalised Sensors for Metal Ions. Polym. Chem. 2012, 3 (11), 3166– 3174, DOI: 10.1039/c2py20350bGoogle Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVGisbvJ&md5=546ada39794d7cbb342ec917911bab5fSingle chain polymeric nanoparticles as compartmentalised sensors for metal ionsGillissen, Martijn A. J.; Voets, Ilja K.; Meijer, E. W.; Palmans, Anja. R. A.Polymer Chemistry (2012), 3 (11), 3166-3174CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)3,3'-Bis(acylamino)-2,2'-bipyridine substituted benzene-1,3,5-tricarboxamide (BiPy-BTA) grafted polynorbornene polymers were prepd. via ring-opening metathesis polymn. using a third generation Grubbs catalyst. The polymers fold intramolecularly via π-π interactions into fluorescent, compartmentalized particles of nanometer-size in mixts. of THF and methylcyclohexane. Spectroscopic and light scattering techniques show that the compact conformation of the folded polymer is affected by increasing the BiPy-BTA functionalization degree and by changing the solvent polarity. Changes in the conformation are accompanied by changes in the fluorescence intensity. Due to the affinity of the 2,2'-bipyridine units for metal ions such as Cu, the particles obtained are effective sensors for these metals. The compartmentalization of the binding motifs in SCPNs proves to be advantageous in sensor applications of these particles.
- 47Terashima, T.; Sugita, T.; Fukae, K.; Sawamoto, M. Synthesis and Single-Chain Folding of Amphiphilic Random Copolymers in Water. Macromolecules 2014, 47 (2), 589– 600, DOI: 10.1021/ma402355vGoogle Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXnsFersg%253D%253D&md5=05f57b8e32e1a5aa4f9f830e7ff5e3ddSynthesis and Single-Chain Folding of Amphiphilic Random Copolymers in WaterTerashima, Takaya; Sugita, Takanori; Fukae, Kaoru; Sawamoto, MitsuoMacromolecules (Washington, DC, United States) (2014), 47 (2), 589-600CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Amphiphilic random methacrylate copolymers, consisting of poly(ethylene glycol) (PEG) and alkyl pendent groups, undergo reversible single-chain self-folding in water via intramol. hydrophobic interaction, to generate a dynamic unimol. hydrophobic nanospace, similar in shape but structurally different relative to micelles and microgel star polymers. These copolymers were prepd. by the ruthenium-catalyzed living radical copolymn. of a PEG methacrylate (PEGMA) and an alkyl methacrylate (RMA; R, -CnH2n+1, n = 1-18), where copolymer compn., d.p., and hydrophobic R moiety were varied. Detailed structural and chain-folding characterization has revealed: single-chain folding is favored with the RMA content 20-40 mol % per chain; the hydrophobic inner compartment (or the self-folded structure) is stable even at a high polymer concn. (up to ∼6 wt. %); and folded-unfolded transition occurs on addn. of methanol or by elevating soln. temp., finally to phase-sepn. above a lower crit. soln. temp.
- 48Terashima, T.; Mes, T.; De Greef, T. F. A.; Gillissen, M. A. J.; Besenius, P.; Palmans, A. R. A.; Meijer, E. W. Single-Chain Folding of Polymers for Catalytic Systems in Water. J. Am. Chem. Soc. 2011, 133 (13), 4742– 4745, DOI: 10.1021/ja2004494Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjtFeku70%253D&md5=a4f7c01b74384b16c8f00c7b2ab04641Single-Chain Folding of Polymers for Catalytic Systems in WaterTerashima, Takaya; Mes, Tristan; De Greef, Tom F. A.; Gillissen, Martijn A. J.; Besenius, Pol; Palmans, Anja R. A.; Meijer, E. W.Journal of the American Chemical Society (2011), 133 (13), 4742-4745CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Enzymes are a source of inspiration for chemists attempting to create versatile synthetic catalysts. In order to arrive at a polymeric chain carrying catalytic units sepd. spatially, it is a prerequisite to fold these polymers in water into well-defined compartmentalized architectures thus creating a catalytic core. Herein, we report the synthesis, phys. properties, and catalytic activity of a water-sol. segmented terpolymer in which a helical structure in the apolar core is created around a ruthenium-based catalyst. The supramol. chirality of this catalytic system is the result of the self-assembly of benzene-1,3,5-tricarboxamide side chains, while the catalyst arises from the sequential ruthenium-catalyzed living radical polymn. of the different monomers followed by ligand exchange. The polymers exhibit a two-state folding process and show transfer hydrogenation in water.
- 49Matsumoto, M.; Sawamoto, M.; Terashima, T. Orthogonal Folding of Amphiphilic/Fluorous Random Block Copolymers for Double and Multicompartment Micelles in Water. ACS Macro Lett. 2019, 8 (3), 320– 325, DOI: 10.1021/acsmacrolett.9b00078Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXkt1eisb4%253D&md5=bf44aff6acc6c9b7b9f35cbf970b1d9fOrthogonal Folding of Amphiphilic/Fluorous Random Block Copolymers for Double and Multicompartment Micelles in WaterMatsumoto, Mayuko; Sawamoto, Mitsuo; Terashima, TakayaACS Macro Letters (2019), 8 (3), 320-325CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)Here, we report orthogonal folding and self-assembly systems of amphiphilic/fluorous random block copolymers for double core and multicompartment micelles in water. For this, we developed the precision folding techniques of polymer chains via the selective self-assembly of the pendant groups. Typically, A/C-B/C random block copolymers were designed: Hydrophobic dodecyl groups (A) and fluorous fluorinated octyl groups (B) were introduced into the resp. blocks, while hydrophilic poly(ethylene glycol) chains (C) were randomly incorporated into all the segments. By controlling the chain length and compn. of the resp. blocks, the copolymers induce orthogonal single-chain folding in water to form double-compartment micelles comprising hydrophobic and fluorous cores. The copolymers were site-selectively folded in a fluoroalc. to result in tadpole unimer micelles comprising a hydrophobic A/C unimer micelle and an unfolded fluorous B/C chain. Addnl., asym. A/C-B/C random block copolymers with short and highly hydrophobic or fluorous segments were effective for multicompartment micelles in water.
- 50Matsumoto, M.; Terashima, T.; Matsumoto, K.; Takenaka, M.; Sawamoto, M. Compartmentalization Technologies via Self-Assembly and Cross-Linking of Amphiphilic Random Block Copolymers in Water. J. Am. Chem. Soc. 2017, 139 (21), 7164– 7167, DOI: 10.1021/jacs.7b03152Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXnslKis7k%253D&md5=5de60e0783e388ccc071895904f7af55Compartmentalization Technologies via Self-Assembly and Cross-Linking of Amphiphilic Random Block Copolymers in WaterMatsumoto, Mayuko; Terashima, Takaya; Matsumoto, Kazuma; Takenaka, Mikihito; Sawamoto, MitsuoJournal of the American Chemical Society (2017), 139 (21), 7164-7167CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Orthogonal self-assembly and intramol. crosslinking of amphiphilic random block copolymers in water afforded an approach to tailor-make well-defined compartments and domains in single polymer chains and nanoaggregates. For a double compartment single-chain polymer, an amphiphilic random block copolymer bearing hydrophilic poly(ethylene glycol) (PEG) and hydrophobic dodecyl, benzyl, and olefin pendants was synthesized by living radical polymn. (LRP) and postfunctionalization; the dodecyl and benzyl units were incorporated into the different block segments, whereas PEG pendants were statistically attached along a chain. The copolymer self-folded via the orthogonal self-assembly of hydrophobic dodecyl and benzyl pendants in water, followed by intramol. crosslinking, to form a single-chain polymer carrying double yet distinct hydrophobic nanocompartments. A single-chain crosslinked polymer with a chlorine terminal served as a globular macroinitiator for LRP to provide an amphiphilic tadpole macromol. comprising a hydrophilic nanoparticle and a hydrophobic polymer tail; the tadpole thus self-assembled into multicompartment aggregates in water.
- 51Neal, T. J.; Beattie, D. L.; Byard, S. J.; Smith, G. N.; Murray, M. W.; Williams, N. S. J.; Emmett, S. N.; Armes, S. P.; Spain, S. G.; Mykhaylyk, O. O. Self-Assembly of Amphiphilic Statistical Copolymers and Their Aqueous Rheological Properties. Macromolecules 2018, 51 (4), 1474– 1487, DOI: 10.1021/acs.macromol.7b02134Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitlarsrw%253D&md5=a06b2c97ac88e4a06405b09b3e1dcdd3Self-Assembly of Amphiphilic Statistical Copolymers and Their Aqueous Rheological PropertiesNeal, Thomas J.; Beattie, Deborah L.; Byard, Sarah J.; Smith, Gregory N.; Murray, Martin W.; Williams, Neal S. J.; Emmett, Simon N.; Armes, Steven P.; Spain, Sebastian G.; Mykhaylyk, Oleksandr O.Macromolecules (Washington, DC, United States) (2018), 51 (4), 1474-1487CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A range of poly(Bu methacrylate-stat-methacrylic acid) [P(BMA-stat-MAA)] statistical copolymers of various compns. and mol. wts. ranging from 5 to 30 kDa were prepd. using either reversible addn.-fragmentation chain transfer (RAFT) soln. copolymn. or conventional free radical polymn. in isopropanol (IPA). On diln. with water, these amphiphilic copolymers self-assembled to form spherical nano-objects as confirmed by small-angle X-ray scattering (SAXS) and transmission electron microscopy. Various structural models were examd. to ext. information regarding the mean nano-object size and morphol. It is found that nano-object radii are independent of copolymer mol. wt. but depend on the copolymer compn.: the smaller the amt. of MAA units in the copolymer chains, the larger the nano-objects that are formed. Combined SAXS and aq. electrophoretic measurements indicated that most of the MAA units are located at the nano-object surface. Furthermore, SAXS and rheol. measurements were used to monitor the effect of solvent compn. on the copolymer morphol. both at a fixed copolymer concn. (either 1 or 25 wt. %) and also for a gradual variation in copolymer concns. (from 1 to 40 wt. %) when adding water to the initial copolymer soln. in IPA. These studies revealed that the copolymers are present in soln. as molecularly dissolved Gaussian chains when the solvent compn. is IPA-rich. However, the copolymer chains self-assemble into spherical nano-objects when the solvent compn. is water-rich. At intermediate solvent compns., SAXS anal. confirmed the formation of an interconnected nano-object network, which accounts for the apparently anomalous increase in soln. viscosity on diln. indicated by rheol. measurements.
- 52Hattori, G.; Takenaka, M.; Sawamoto, M.; Terashima, T. Nanostructured Materials via the Pendant Self-Assembly of Amphiphilic Crystalline Random Copolymers. J. Am. Chem. Soc. 2018, 140 (27), 8376– 8379, DOI: 10.1021/jacs.8b03838Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtFCms77J&md5=06ffde6a8667e2617a6a6ab1b0aabcdeNanostructured Materials via the Pendant Self-Assembly of Amphiphilic Crystalline Random CopolymersHattori, Goki; Takenaka, Mikihito; Sawamoto, Mitsuo; Terashima, TakayaJournal of the American Chemical Society (2018), 140 (27), 8376-8379CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Versatile self-assembly systems to nanostructured materials in both solid and soln. were developed with common amphiphilic random copolymers bearing hydrophilic poly(ethylene glycol) (PEG) and hydrophobic cryst. octadecyl pendants. The copolymers efficiently induced precision self-assembly of the pendants to provide not only core-cryst., thermoresponsive micelles and vesicles in water and reverse micelles in hexane but also sub-10 nm lamellar or spherical microphase sepn. structure in solid. Typically, the solid random copolymers with 50-80 mol % octadecyl units formed lamellar structure of a hydrophilic PEG layer and a hydrophobic, cryst. octadecyl layer. Importantly, the domain spacing is about 5 nm, much smaller than that generally obtained with conventional block copolymers. The domain structure is controlled by compn., independent of chain length. The copolymers further gave various thermoresponsive, compartmentalized materials in aq. and org. media, where the 3D structure can be also controlled by the compn. and sample prepn. protocols.
- 53Ilavsky, J.; Jemian, P. R. Irena : Tool Suite for Modeling and Analysis of Small-Angle Scattering. J. Appl. Crystallogr. 2009, 42 (2), 347– 353, DOI: 10.1107/S0021889809002222Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXjsFSnsbY%253D&md5=ebf014363d395cb4180fe0fd96f7f714Irena: tool suite for modeling and analysis of small-angle scatteringIlavsky, Jan; Jemian, Peter R.Journal of Applied Crystallography (2009), 42 (2), 347-353CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Irena, a tool suite for anal. of both x-ray and neutron small-angle scattering (SAS) data within the com. Igor Pro application, brings together a comprehensive suite of tools useful for studies in materials science, physics, chem., polymer science and other fields. In addn. to Guinier and Porod fits, the suite combines a variety of advanced SAS data evaluation tools for the modeling of size distribution in the dil. limit using max. entropy and other methods, dil. limit small-angle scattering from multiple noninteracting populations of scatterers, the pair-distance distribution function, a unified fit, the Debye-Bueche model, the reflectivity (x-ray and neutron) using Parratt's formalism, and small-angle diffraction. There are also a no. of support tools, such as a data import/export tool supporting a broad sampling of common data formats, a data modification tool, a presentation-quality graphics tool optimized for small-angle scattering data, and a neutron and x-ray scattering contrast calculator. These tools are brought together into one suite with consistent interfaces and functionality. The suite allows robust automated note recording and saving of parameters during export.
- 54Horcas, I.; Fernández, R.; Gómez-Rodríguez, J. M.; Colchero, J.; Gómez-Herrero, J.; Baro, A. M. WSXM: A Software for Scanning Probe Microscopy and a Tool for Nanotechnology. Rev. Sci. Instrum. 2007, 78, 013705, DOI: 10.1063/1.2432410Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXis1Srsrg%253D&md5=ff73289035c4652ebb9800e362892410WSXM: a software for scanning probe microscopy and a tool for nanotechnologyHorcas, I.; Fernandez, R.; Gomez-Rodriguez, J. M.; Colchero, J.; Gomez-Herrero, J.; Baro, A. M.Review of Scientific Instruments (2007), 78 (1), 013705/1-013705/8CODEN: RSINAK; ISSN:0034-6748. (American Institute of Physics)In this work we briefly describe the most relevant features of WSXM, a freeware scanning probe microscopy software based on MS-Windows. The article is structured in three different sections: The introduction is a perspective on the importance of software on scanning probe microscopy. The second section is devoted to describe the general structure of the application; in this section the capabilities of WSXM to read third party files are stressed. Finally, a detailed discussion of some relevant procedures of the software is carried out.
- 55Kheshgi, S.; Scriven, L. E. Dewetting: Nucleation and Growth of Dry Regions. Chem. Eng. Sci. 1991, 46 (2), 519– 526, DOI: 10.1016/0009-2509(91)80012-NGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhtVGjs7k%253D&md5=bd33a4450b0cef9f1c03355fbd9dbf60Dewetting: nucleation and growth of dry regionsKheshgi, Haroon S.; Scriven, L. E.Chemical Engineering Science (1991), 46 (2), 519-26CODEN: CESCAC; ISSN:0009-2509.Dewetting of a solid surface covered by a film of nonwetting liq. proceeds from a preexisting dry patch or edge or initiates from some film-thinning disturbance that grows until the film ruptures. Local thinning can be caused by evapn.; by drainage due to gravity or capillarity-driven flow, esp. from sharp surfaces; or by surface tension gradients, such as are caused by surfactants delivered by particles falling on the film. Once a nonwetting film is sufficiently thinned, conjoining (neg. disjoining) pressure can accelerate thinning until rupture. This catastrophic rupture is modeled by solving the Navier-Stokes system approximated for thickness variations over distances that are long compared with the mean film thickness, and augmented with conjoining pressure. Rupture leads to film retraction and formation of a dry patch. These phenomena are visualized via moire topog. Of special interest are local spreading disturbances, where airborne particles fall on the film surface; craters or dry patches often nucleate. Implications for coating operations are discussed.
- 56Xue, L.; Han, Y. Inhibition of Dewetting of Thin Polymer Films. Prog. Mater. Sci. 2012, 57 (6), 947– 979, DOI: 10.1016/j.pmatsci.2012.01.003Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xls1Siu70%253D&md5=0c23ab0ed2e6414db3bffb2e3b764a71Inhibition of dewetting of thin polymer filmsXue, Longjian; Han, YanchunProgress in Materials Science (2012), 57 (6), 947-979CODEN: PRMSAQ; ISSN:0079-6425. (Elsevier Ltd.)A review. Strategies for the inhibition of dewetting of thin polymer films are reviewed. First, a brief introduction to the theory and the dynamics of dewetting of thin polymer films is given. Methods for the inhibition of dewetting of thin polymer films, including the modification of the substrate and the polymer, the crosslinking of the polymer and the addn. of an additive as well as their mechanisms for inhibition are discussed. The chem. modification of the substrate or the polymer itself and the phase sepn. increase the thermodn. stability of the system, while the crosslinking of the polymers reduces the mobility of the polymer chains, kinetically arresting the dewetting. The addn. of an additive appears to thermodynamically and kinetically inhibit the dewetting. Finally, the future outlook in this research field is highlighted.
- 57Barsbay, M.; Güven, O.; Stenzel, M. H.; Davis, T. P.; Barner-Kowollik, C.; Barner, L. Verification of Controlled Grafting of Styrene from Cellulose via Radiation-Induced RAFT Polymerization. Macromolecules 2007, 40 (20), 7140– 7147, DOI: 10.1021/ma070825uGoogle Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtVSgs7bF&md5=7686cbcffa76684773b4600319dc3594Verification of Controlled Grafting of Styrene from Cellulose via Radiation-Induced RAFT PolymerizationBarsbay, Murat; Gueven, Olgun; Stenzel, Martina H.; Davis, Thomas P.; Barner-Kowollik, Christopher; Barner, LeonieMacromolecules (Washington, DC, United States) (2007), 40 (20), 7140-7147CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Reversible addn.-fragmentation chain transfer (RAFT) polymn. was applied to radiation-induced graft polymn. of styrene from cellulose. The grafting of styrene from cellulose substrates using the chain transfer agent cumyl phenyldithioacetate was confirmed by Raman and XPS, differential scanning calorimetery, thermogravimetric anal., SEM, and contact angle anal. Grafted polystyrene chains were cleaved from the cellulose surface by acidic hydrolysis of the cellulose. The no.-av. mol. wt. and polydispersity index of the grafted and the free (nongrafted) polystyrenes obtained under identical conditions were detd. by size exclusion chromatog. Grafted and nongrafted polystyrenes have almost the same (near theor.) mol. wt. and narrow polydispersity, thus proving for the first time the control of the grafting process mediated via RAFT without any prior functionalization of the surface.
- 58Eisenberg, A.; Yokoyama, T.; Sambalido, E. Dehydration Kinetics and Glass Transition of Poly(Acrylic Acid). J. Polym. Sci. Part A-1 Polym. Chem. 1969, 7 (7), 1717– 1728, DOI: 10.1002/pol.1969.150070714Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXksVartQ%253D%253D&md5=7168b32a3f0f9c9b61cb323407e90d32Dehydration kinetics and glass transition of poly(acrylic acid)Eisenberg, Adi; Yokoyama, Tetsuo; Sambalido, EmmaJournal of Polymer Science, Part A-1: Polymer Chemistry (1969), 7 (7), 1717-28CODEN: JPSPC3; ISSN:0449-296X.The kinetics of dehydration and decarboxylation as well as the glass transition temp. as a function of anhydride content were measured for poly(acrylic acid) (I). The glass transition of I is of the order of 103° and increases with increasing anhydride content, reaching an extrapolated value of 140° for the pure linear anhydride. Anhydride formation is a first-order reaction, as is also decarboxylation, the latter being much slower than the former. The rate consts. are: for dehydration, ka = 2.5 × 109 exp (-26,000/RT); for decarboxylation, kd = 2.9 × 108 exp (-27,000/RT). Anhydride formation occurs primarily by an intramol. process.
- 59Chapin, E. C.; Ham, G. E.; Mills, C. L. Copolymerization. VII. Relative Rates of Addition of Various Monomers in Copolymerization. J. Polym. Sci. 1949, 4, 597– 604, DOI: 10.1002/pol.1949.120040505Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG3cXptVOn&md5=e78bf52fb289122c0952eccccb62a5efCopolymerization. VII. Relative rates of addition of various monomers in copolymerizationChapin, Earl C.; Ham, George E.; Mills, Charles L.Journal of Polymer Science (1949), 4 (), 597-604CODEN: JPSCAU; ISSN:0022-3832.cf. C.A. 43, 8736g; 42, 4151i. Data are presented on monomer reactivity in copolymerization for these 13 systems: styrene (I)-methacrylic acid (II), I-acrylic acid (III), vinylidene chloride (IV)-crotonic acid (V), vinyl acetate (VI)-V, acrylonitrile (VII)-vinyl 2-ethylhexoate (VIII), VII-vinyl formate (IX), VII-vinyl benzoate (X), VII-α-acetoxystyrene (XI), VII-allyl chloride (XII), IV-X, IV-methyl isopropenyl ketone (XIII), vinyl chloride (XIV)-vinyl isobutyl ether (XV), and I-citraconic anydride (XVI). The reactivity ratios, γ1 and γ2, for these copolymerizations were obtained by choosing values which gave the best curve fit on substitution in the differential copolymer equation. The azeotrope of the system I-II occurs at 74% II while that of I-III occurs at 48% III. By using the Price and Alfrey relationship for the consts. Q and e these values are interpreted as evidence of higher resonance stability of the radical adduct formed from II than from III and lower polarity of II than III. Azeotropes also occur in the systems VI-V, VII-XI, and I-XVI. The copolymerization of I and XVI yielded a copolymerization curve similar to that of I-maleic anhydride reported by Alfrey and Lavin (C.A. 40, 1149.2); evidently XVI is less reactive toward a I-free radical than maleic anhydride.
- 60Mayadunne, R. T. A.; Rizzardo, E.; Chiefari, J.; Krstina, J.; Moad, G.; Postma, A.; Thang, S. H. Radical Polymerization in Two Steps. Macromolecules 2000, 33, 243– 245, DOI: 10.1021/ma991451aGoogle Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhtlWgtA%253D%253D&md5=a0ad88898f76e3c4b5eca06698d39614Living polymers by the use of trithiocarbonates as reversible addition-fragmentation chain transfer (RAFT) agents. ABA triblock copolymers by radical polymerization in two stepsMayadunne, Roshan T. A.; Rizzardo, Ezio; Chiefari, John; Krstina, Julia; Moad, Graeme; Postma, Almar; Thang, San H.Macromolecules (2000), 33 (2), 243-245CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Trithiocarbonates RS-CS-SR', where R and R' are homolytic leaving groups, were used as effective reversible addn.-fragmentation chain transfer agents. Styrene, Me acrylate, Bu acrylate and Me methacrylate were polymd. in the presence of various trithiocarbonate compds. (R/R': Me/C(Me)2CN, Me/CH(Ph)COOH, benzyl/benzyl, and α-methylbenzyl/α-methylbenzyl) and mol. wt. and conversion data were detd. Homopolymers and block copolymers of controlled mol. wt. and narrow polydispersity were prepd. Important ABA triblock copolymers of predetd. mol. wt. are accessible in 2 steps.
- 61Buzin, A. I.; Pyda, M.; Costanzo, P.; Matyjaszewski, K.; Wunderlich, B. Calorimetric Study of Block-Copolymers of Poly(n-Butyl Acrylate) and Gradient Poly(n-Butyl Acrylate-Co-Methyl Methacrylate). Polymer (Guildf). 2002, 43 (20), 5563– 5569, DOI: 10.1016/S0032-3861(02)00358-0Google ScholarThere is no corresponding record for this reference.
- 62Sanchez-Fernandez, A.; Edler, K. J.; Arnold, T.; Heenan, R. K.; Porcar, L.; Terrill, N. J.; Terry, A. E.; Jackson, A. J. Micelle Structure in a Deep Eutectic Solvent: A Small-Angle Scattering Study. Phys. Chem. Chem. Phys. 2016, 18 (20), 14063– 14073, DOI: 10.1039/C6CP01757FGoogle Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XmvFOlsLc%253D&md5=a46225f5c8786103c0dcef785859f328Micelle structure in a deep eutectic solvent: a small-angle scattering studySanchez-Fernandez, A.; Edler, K. J.; Arnold, T.; Heenan, R. K.; Porcar, L.; Terrill, N. J.; Terry, A. E.; Jackson, A. J.Physical Chemistry Chemical Physics (2016), 18 (20), 14063-14073CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)In recent years many studies into green solvents have been undertaken and deep eutectic solvents (DES) have emerged as sustainable and green alternatives to conventional solvents since they may be formed from cheap non-toxic org. precursors. In this study we examine amphiphile behavior in these novel media to test our understanding of amphiphile self-assembly within environments that have an intermediate polarity between polar and non-polar extremes. We have built on our recently published results to present a more detailed structural characterization of micelles of sodium dodecylsulfate (SDS) within the eutectic mixt. of choline chloride and urea. Here we show that SDS adopts an unusual cylindrical aggregate morphol., unlike that seen in water and other polar solvents. A new morphol. transition to shorter aggregates was found with increasing concn. The self-assembly of SDS was also investigated in the presence of water; which promotes the formation of shorter aggregates.
- 63Gaillard, N.; Guyot, A.; Claverie, J. Block Copolymers of Acrylic Acid and Butyl Acrylate Prepared by Reversible Addition-Fragmentation Chain Transfer Polymerization: Synthesis, Characterization, and Use in Emulsion Polymerization. J. Polym. Sci. Part A Polym. Chem. 2003, 41 (5), 684– 698, DOI: 10.1002/pola.10606Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhtlCnur0%253D&md5=d40c44d67105fc9ec9ea57155237b4feBlock copolymers of acrylic acid and butyl acrylate prepared by reversible addition-fragmentation chain transfer polymerization: synthesis, characterization, and use in emulsion polymerizationGaillard, Nicolas; Guyot, Alain; Claverie, JeromeJournal of Polymer Science, Part A: Polymer Chemistry (2003), 41 (5), 684-698CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)Amphiphilic block copolymers of poly(acrylic acid-b-Bu acrylate) (I) were prepd. by reversible addn.-fragmentation chain transfer polymn. in a one-pot reaction. I were characterized by NMR, static and dynamic light scattering, tensiometry, and size exclusion chromatog. The aggregation characteristics of I corresponded to those theor. predicted for a star micelle. In a Bu acrylate-Me methacrylate emulsion copolymn., low amts. of I could stabilize the latexes with solid contents up to 50%.
- 64Chernikova, E. V.; Lysenko, E. A.; Serkhacheva, N. S.; Prokopov, N. I. Self-Assembly of Amphiphilic Block Copolymers during Reversible Addition-Fragmentation Chain Transfer Heterophase Polymerization: Problems, Achievements, and Outlook. Polym. Sci. - Ser. C 2018, 60 (2), 192– 218, DOI: 10.1134/S1811238218020042Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitlKmtbnI&md5=06f47b161ab97bcc2458ee70e63796dfSelf-Assembly of Amphiphilic Block Copolymers during Reversible Addition-Fragmentation Chain Transfer Heterophase Polymerization: Problems, Achievements, and OutlookChernikova, E. V.; Lysenko, E. A.; Serkhacheva, N. S.; Prokopov, N. I.Polymer Science, Series C: Selected Topics (2018), 60 (S1), 192-218CODEN: PSSCCH; ISSN:1555-614X. (Pleiades Publishing, Ltd.)The formation of dispersions of amphiphilic block copolymer particles with the controlled morphol. via heterophase polymn. mediated by reversible addn.-fragmentation agents is considered. Variants of dispersion, emulsion, and seeded polymns. are analyzed, and the mechanism of this process is discussed. Special attention is focused on issues related to control over the morphol. of the formed particles immediately during synthesis and methods of its transformation in the resulting dispersions.
- 65Chenal, M.; Rieger, J.; Vechambre, C.; Chenal, J.-M.; Chazeau, L.; Creton, C.; Bouteiller, L. Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard Phase. Macromol. Rapid Commun. 2013, 34, 1524– 1529, DOI: 10.1002/marc.201300410Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlahtLbO&md5=4d4bc64299888278198f8cdf39b17d57Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard PhaseChenal, Marion; Rieger, Jutta; Vechambre, Cyril; Chenal, Jean-Marc; Chazeau, Laurent; Creton, Costantino; Bouteiller, LaurentMacromolecular Rapid Communications (2013), 34 (19), 1524-1529CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)In this study, aq. emulsion polymn. of Bu acrylate is performed in batch conditions without surfactants using a poly(acrylic acid)-trithiocarbonate macro-RAFT agent to control the polymn. and to stabilize the emulsion. According to the polymn.-induced self-assembly (PISA) approach, well-defined amphiphilic PAA-b-PBA diblock copolymers form and self-assemble during synthesis to yield highly stable core-shell particles with an extremely thin hard PAA shell. For the first time, we report here the specific properties of films obtained from these particular latexes. After drying the aq. dispersion, tough and transparent films are obtained. Although the films are not chem. crosslinked, they do not dissolve in good solvents for PBA. Moreover, they remain transparent even after immersion in water. Rheol. shows that the films are both stiff and ductile, thanks to the nanostructured but very low vol. fraction (less than 3 wt%) of PAA forming a percolating network in the soft PBA. Compared with conventional core-shell-based films, this approach affords for the first time a route to a thin percolating honeycomb nanostructure with a sharp and strong interface between the two phases. The versatility of the synthetic procedure opens perspectives for a large range of functional materials.
- 66Pedersen, J. S.; Gerstenberg, M. C. Scattering Form Factor of Block Copolymer Micelles. Macromolecules 1996, 29 (4), 1363– 1365, DOI: 10.1021/ma9512115Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XltlWjsA%253D%253D&md5=e6f549fc773b455a50c0d6e157ca4981Scattering Form Factor of Block Copolymer MicellesPedersen, Jan Skov; Gerstenberg, Michael C.Macromolecules (1996), 29 (4), 1363-5CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The form factor of a micelle model with a spherical core and Gaussian polymer chains attached to the surface was calcd. anal. and the results were compared to Monte Carlo simulations. Excluded vol. interactions between the core and the polymers were introduced in the simulations. The expansion of the coils due to this effect can be mimicked in the anal. calcns. by moving the center of mass of the chains away from the surface of the core. The anal. expression for the form factor has been used for analyzing small-angle scattering data.
- 67Balmer, J. A.; Mykhaylyk, O. O.; Schmid, A.; Armes, S. P.; Fairclough, J. P. A.; Ryan, A. J. Characterization of Polymer-Silica Nanocomposite Particles with Core-Shell Morphologies Using Monte Carlo Simulations and Small Angle X-Ray Scattering. Langmuir 2011, 27 (13), 8075– 8089, DOI: 10.1021/la201319hGoogle Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXntlCqtrg%253D&md5=9ea3ca1b4a8edcc48f631a3152519b91Characterization of Polymer-Silica Nanocomposite Particles with Core-Shell Morphologies using Monte Carlo Simulations and Small Angle X-ray ScatteringBalmer, Jennifer A.; Mykhaylyk, Oleksandr O.; Schmid, Andreas; Armes, Steven P.; Fairclough, J. Patrick A.; Ryan, Anthony J.Langmuir (2011), 27 (13), 8075-8089CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A two-population model based on std. small-angle x-ray scattering (SAXS) equations is verified for the anal. of core-shell structures comprising spherical colloidal particles with particulate shells. First, Monte Carlo simulations of core-shell structures are performed to demonstrate the applicability of the model. Three possible shell packings are considered: ordered silica shells due to either charge-dependent repulsive or size-dependent Lennard-Jones interactions or randomly arranged silica particles. In most cases, the two-population model produces an excellent fit to calcd. SAXS patterns for the simulated core-shell structures, together with a good correlation between the fitting parameters and structural parameters used for the simulation. The limits of application are discussed, and then, this two-population model is applied to the anal. of well-defined core-shell vinyl polymer/silica nanocomposite particles, where the shell comprises a monolayer of spherical silica nanoparticles. Comprehensive SAXS anal. of poly(styrene-co-Bu acrylate)/silica colloidal nanocomposite particles (prepd. by the in situ emulsion copolymn. of styrene and Bu acrylate in the presence of a glycerol-functionalized silica sol) allows the overall core-shell particle diam., the copolymer latex core diam. and polydispersity, the mean silica shell thickness, the mean silica diam. and polydispersity, the vol. fractions of the two components, the silica packing d., and the silica shell structure to be obtained. These exptl. SAXS results are consistent with electron microscopy, dynamic light scattering, TG, helium pycnometry, and BET surface area studies. The high electron d. contrast between the (co)polymer and the silica components, together with the relatively low polydispersity of these core-shell nanocomposite particles, makes SAXS ideally suited for the characterization of this system. Also, these results can be generalized for other types of core-shell colloidal particles.
- 68Matsen, M. W.; Thompson, R. B. Equilibrium Behavior of Asymmetric ABA Triblock Copolymer Melts. J. Chem. Phys. 1999, 111, 7139, DOI: 10.1063/1.480006Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXmt1Oisr0%253D&md5=9e964439283738bcd88497f688856f2dEquilibrium behavior of symmetric ABA triblock copolymer meltsMatsen, M. W.; Thompson, R. B.Journal of Chemical Physics (1999), 111 (15), 7139-7146CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Melts of ABA triblock copolymer mols. with identical end blocks are examd. using SCF theory. Phase diagrams are calcd. and compared with those of homologous AB diblock copolymers formed by snipping the triblocks in half. This creates addnl. end segments which decreases the degree of segregation. Consequently, triblock melts remain ordered to higher temps. than their diblock counterparts. We also find that middle-block domains are easier to stretch than end-block domains. As a result, domain spacings are slightly larger, the complex phase regions are shifted towards smaller A-segment compns., and the perforated-lamellar phase becomes more metastable in triblock melts as compared to diblock melts. Although triblock and diblock melts exhibit very similar phase behavior, their mech. properties can differ substantially due to triblock copolymers that bridge between otherwise disconnected A domains. We evaluate the bridging fraction for lamellar, cylindrical, and spherical morphologies to be about 40%-45%, 60%-65%, and 75%-80%, resp. These fractions only depend weakly on the degree of segregation and the copolymer compn.
- 69Ryan, A. J.; Mai, S. M.; Fairclough, J. P. A.; Hamley, I. W.; Booth, C. Ordered Melts of Block Copolymers of Ethylene Oxide and 1,2-Butylene Oxide. Phys. Chem. Chem. Phys. 2001, 3 (15), 2961– 2971, DOI: 10.1039/b102784kGoogle Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlt1ygu74%253D&md5=be0263f79a0e3ef9cf50ea3c3e86b3abOrdered melts of block copolymers of ethylene oxide and 1,2-butylene oxideRyan, Anthony J.; Mai, Shao-Min; Fairclough, J. Patrick A.; Hamley, Ian W.; Booth, ColinPhysical Chemistry Chemical Physics (2001), 3 (15), 2961-2971CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)An account is presented of recent work on specially synthesized diblock, triblock and cyclic block copolymers of ethylene oxide and 1,2-butylene oxide. Simultaneous small-angle X-ray scattering and differential scanning calorimetry have been used to investigate the effects of block architecture on the stabilities and structures of microphase-sepd. melts. Stable lamellar, hexagonal, body-centered cubic and gyroid phases were detected. Phase diagrams are compared, one with another and with those predicted by the exact self-consistent mean-field theory, and center-block stretching in lamellar phases is confirmed.
- 70Mai, Y.; Eisenberg, A. Self-Assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41 (18), 5969, DOI: 10.1039/c2cs35115cGoogle Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1aqsbvL&md5=37964bf5011f9bc7a6aa42c4b612cf91Self-assembly of block copolymersMai, Yiyong; Eisenberg, AdiChemical Society Reviews (2012), 41 (18), 5969-5985CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Block copolymer (BCP) self-assembly has attracted considerable attention for many decades. The present tutorial review introduces the primary principles of BCP self-assembly in bulk and in soln., by describing expts., theories, accessible morphologies, etc.
- 71Lynd, N. A.; Meuler, A. J.; Hillmyer, M. A. Polydispersity and Block Copolymer Self-Assembly. Prog. Polym. Sci. 2008, 33 (9), 875– 893, DOI: 10.1016/j.progpolymsci.2008.07.003Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1Klt77P&md5=30b678d94f8baaad00be1d8bec1759e0Polydispersity and block copolymer self-assemblyLynd, Nathaniel A.; Meuler, Adam J.; Hillmyer, Marc A.Progress in Polymer Science (2008), 33 (9), 875-893CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. Block copolymers consist of two or more chem. distinct polymers that are covalently bound. These materials self-assemble into fascinating mesostructures with features on the nanometer length scale and have been the subject of intense research interest for about four decades. These efforts have generally focused on model block copolymer systems where the mol. wt. distributions of all blocks are very narrow. Traditionally, many block copolymer systems have been prepd. by living anionic polymn. and thus usually exhibit narrow mol. wt. distributions in all blocks. Therefore, the assumption of monodisperse blocks that greatly simplifies theor. work is on solid exptl. ground. Prepn. of block copolymers with relatively broad mol. wt. distributions in one or more block has become increasingly common, however, as use of synthetic techniques such as controlled radical polymn. has proliferated. A review. Advances in these techniques have increased the no. of monomers readily incorporated into block copolymers and potentially will drive com. costs down. These polymn. strategies often, however, result in broader mol. wt. distributions than are typically obtained using living anionic, cationic, or metal-catalyzed techniques; understanding polydispersity effects should aid deployment of these block copolymers in advanced materials applications. This review describes both theor. and exptl. investigations of the effects of polydispersity on the melt-phase morphol. behavior of block copolymers. The summary includes research efforts focused on both continuous mol. wt. distributions and multicomponent blends. The review concludes with a summary and outlook on the potential utility of polydispersity as a tool to tune the morphol. behavior of block copolymers.
- 72Stieger, M.; Pedersen, J. S.; Lindner, P.; Richtering, W. Are Thermoresponsive Microgels Model Systems for Concentrated Colloidal Suspensions? A Rheology and Small-Angle Neutron Scattering Study. Langmuir 2004, 20 (17), 7283– 7292, DOI: 10.1021/la049518xGoogle Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXls1Kkt7w%253D&md5=16b72e479d1666a9e50320ada43e0381Are Thermoresponsive Microgels Model Systems for Concentrated Colloidal Suspensions? A Rheology and Small-Angle Neutron Scattering StudyStieger, Markus; Pedersen, Jan Skov; Lindner, Peter; Richtering, WalterLangmuir (2004), 20 (17), 7283-7292CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The structure of concd. temp.-sensitive poly(N-isopropylacrylamide) (PNiPAM) microgel suspensions has been investigated employing rheol. and small-angle neutron scattering (SANS). A previously described model expression for the particle form factor Pinho(q) is extended by a model hard sphere structure factor S(q), and the av. radial d. profiles φ(r) are calcd. from the amplitude of the form factor A(q) and the structure factor S(q). By this procedure, a direct real space description of the spatial ordering in the neighborhood of a single particle is obtained. The overall particle size and the correlation length ξ of the concn. fluctuations of the internal polymer network decrease with concn., revealing the increasing compression of the spheres. Thus, the particle form factor Pinho(q) of the swollen PNiPAM microgels depends on concn. The particle-particle interaction potential does not change significantly between 25 and 32 °C. Even approx. 1 K below the lower crit. soln. temp. (LCST), the exptl. scattering intensity distributions I(q)/c are described very well by the hard sphere structure factor when an equiv. hard sphere particle size RHS and vol. fraction ηHS are used. Microgels with different degrees of crosslinking and particle size resemble true hard sphere behavior up to effective vol. fractions of φeff < 0.35. At higher effective vol. fractions φeff > 0.35 strong deviations from true hard spheres are obsd. Interpenetration of the outer, less crosslinked regions of the soft spheres as well as particle compression occurred at higher concns. In agreement with this, the equil. colloidal phase behavior and rheol. also has some features of soft sphere systems. At temps. well above the LCST, the interaction potential becomes strongly attractive and the collapsed microgel spheres form aggregates consisting of flocculated particles without significant long-range order. Hence, an attractive interaction potential in concd. suspensions of PNiPAM microgels leads to distinctively different structures as compared to attractive hard sphere colloids. When the peculiar structural properties of the PNiPAM microgels are considered, they can be used as model systems in colloidal science.
- 73De Kruif, C. G.; Briels, W. J.; May, R. P.; Vrij, A. Hard-Sphere Colloidal Silica Dispersions. the Structure Factor Determined with SANS. Langmuir 1988, 4 (3), 668– 676, DOI: 10.1021/la00081a029Google Scholar73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXitFSqs7s%253D&md5=1ac17a1929e4ab37b21bec0554d72a89Hard-sphere colloidal silica dispersions. The structure factor determined with SANSDe Kruif, C. G.; Briels, W. J.; May, R. P.; Vrij, A.Langmuir (1988), 4 (3), 668-76CODEN: LANGD5; ISSN:0743-7463.Small-angle neutron-scattering expts. were performed on the D11 instrument at Grenoble. The colloidal dispersions in cyclohexane studied contained silica particles sterically stabilized by octadecyl chains terminally grafted to the surface. From dil. dispersions the mean particle radius and the radius distribution were obtained. With these single-particle parameters the scattering behavior can be modeled with a hard-sphere interaction potential at all vol. fractions up to .vphi. = 0.4-0.5. Polydispersity effects on the representation of the structure factor were explicitly taken into account. These results confirm and amplify previous results on the modeling of a silica dispersion as a hard-sphere supramol. fluid.
- 74Pedersen, J. S.; Gerstenberg, M. C. Scattering Form Factor of Block Copolymer Micelles. Macromolecules 1996, 29 (4), 1363– 1365, DOI: 10.1021/ma9512115Google Scholar74https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XltlWjsA%253D%253D&md5=e6f549fc773b455a50c0d6e157ca4981Scattering Form Factor of Block Copolymer MicellesPedersen, Jan Skov; Gerstenberg, Michael C.Macromolecules (1996), 29 (4), 1363-5CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The form factor of a micelle model with a spherical core and Gaussian polymer chains attached to the surface was calcd. anal. and the results were compared to Monte Carlo simulations. Excluded vol. interactions between the core and the polymers were introduced in the simulations. The expansion of the coils due to this effect can be mimicked in the anal. calcns. by moving the center of mass of the chains away from the surface of the core. The anal. expression for the form factor has been used for analyzing small-angle scattering data.
- 75Lesage De La Haye, J.; Martin-Fabiani, I.; Schulz, M.; Keddie, J. L.; D’Agosto, F.; Lansalot, M. Hydrophilic MacroRAFT-Mediated Emulsion Polymerization: Synthesis of Latexes for Cross-Linked and Surfactant-Free Films. Macromolecules 2017, 50 (23), 9315– 9328, DOI: 10.1021/acs.macromol.7b01885Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvVWrs7fL&md5=549d77c3d9ebacd5678b9afdaea0f5c7Hydrophilic MacroRAFT-Mediated Emulsion Polymerization: Synthesis of Latexes for Cross-Linked and Surfactant-Free FilmsLesage de la Haye, Jennifer; Martin-Fabiani, Ignacio; Schulz, Malin; Keddie, Joseph L.; D'Agosto, Franck; Lansalot, MurielMacromolecules (Washington, DC, United States) (2017), 50 (23), 9315-9328CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A major drawback of conventional emulsion polymers arises from the presence of migrating low molar mass surfactants that contribute to poor water barrier properties and low adhesion to substrates. In this paper, we demonstrate how living polymer chains obtained by reversible addn.-fragmentation chain transfer (RAFT) can be used as an efficient stabilizer in emulsion polymn., leading to the prodn. of surfactant-free latexes, which then form crosslinked films with beneficial properties. Hydrophilic poly(methacrylic acid) (PMAA) chains obtained by RAFT performed in water are used to mediate emulsion polymn. and produce film-forming latex particles from mixts. of Me methacrylate, Bu acrylate, and styrene. Stable dispersions of particles with sizes between 100 and 200 nm are obtained, with very low amts. of coagulum (<0.5 wt. %). The particles are stabilized by the PMAA segment of amphiphilic block copolymers formed during the polymn. Remarkably, low amts. of PMAA chains (from 1.5 down to 0.75 wt. %) are enough to ensure particle stabilization. Only traces of residual PMAA macroRAFT agents are detected in the final latexes, showing that most of them are successfully chain extended and anchored on the particle surface. The glass transition temp. of the final material is adjusted by the compn. of the hydrophobic monomer mixt. so that film formation occurs at room temp. Conventional crosslinking strategies using addnl. hydrophobic comonomers, such as 1,3-butanediol diacrylate (BuDA), diacetone acrylamide (DAAm), and (2-acetoacetoxy)ethyl methacrylate (AAEM), are successfully applied to these formulations as attested by gel fractions of 100%. When particles are internally crosslinked with BuDA, chain interdiffusion between particles is restricted, and a weak and brittle film is formed. In contrast, when DAAm-contg. chains undergoes crosslinking during film formation, full coalescence is achieved along with the creation of a crosslinked network. The resulting film has a higher Young's modulus and tensile strength as a result of crosslinking. This synthetic strategy advantageously yields a surfactant-free latex that can be formed into a film at room temp. with mech. properties that can be tuned via the crosslinking d.
- 76Martín-Fabiani, I.; Lesage De La Haye, J.; Schulz, M.; Liu, Y.; Lee, M.; Duffy, B.; D’Agosto, F.; Lansalot, M.; Keddie, J. L. Enhanced Water Barrier Properties of Surfactant-Free Polymer Films Obtained by MacroRAFT-Mediated Emulsion Polymerization. ACS Appl. Mater. Interfaces 2018, 10 (13), 11221– 11232, DOI: 10.1021/acsami.8b01040Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXktFSqu7s%253D&md5=eb3947a987c1093495342a2f1973200aEnhanced Water Barrier Properties of Surfactant-Free Polymer Films Obtained by MacroRAFT-Mediated Emulsion PolymerizationMartin-Fabiani, Ignacio; Lesage de la Haye, Jennifer; Schulz, Malin; Liu, Yang; Lee, Michelle; Duffy, Brendan; D'Agosto, Franck; Lansalot, Muriel; Keddie, Joseph L.ACS Applied Materials & Interfaces (2018), 10 (13), 11221-11232CODEN: AAMICK; ISSN:1944-8244. (American Chemical Society)The presence of low-molar-mass surfactants in latex films results in detrimental effects on their water permeability, gloss, and adhesion. For applications such as coatings, there is a need to develop formulations that do not contain surfactants and have better water barrier properties. Having previously reported the synthesis of surfactant-free latex particles in water using low amts. (<2 wt %) of chains synthesized by controlled radical polymn., here we study the water barrier properties of films made from these particles and their application in anticorrosion coatings. When films cast from aq. dispersions of acrylate copolymer particles stabilized with poly(sodium 4-styrenesulfonate) (PSSNa) were immersed in water for 3 days, they sorbed only 4 wt % water. This uptake is only slightly higher than the value predicted for the pure copolymer, indicating that the neg. effects of any particle boundaries and hydrophilic-stabilizing mols. are minimal. This sorption of liq. water is 5 times lower than what is found in films cast from particles stabilized with the same proportion of poly(methacrylic acid) (PMAA), which is more hydrophilic than PSSNa. In water vapor with 90% relative humidity, the PSSNa-based film had an equil. sorption of only 4 wt %. A small increase in the PMAA content has a strong and neg. impact on the barrier properties. NMR relaxometry on polymer films after immersion in water shows that water clusters have the smallest size in the films contg. PSSNa. Furthermore, these films retain their optical clarity during immersion in liq. water for up to 90 min, whereas all other compns. quickly develop opacity ("water whitening") as a result of light scattering from sorbed water. This implies a remarkably complete coalescence and a very small d. of defects, which yields properties matching those of some solvent-borne films. The latex stabilized with PSSNa is implemented as the binder in a paint formulation for application as an anticorrosive barrier coating on steel substrates and evaluated in accelerated weathering and corrosion tests. Our results demonstrate the potential of self-stabilized latex particles for the development of different applications, such as waterborne protective coatings and pressure-sensitive adhesives.
- 77Singh, K. B.; Tirumkudulu, M. S. Cracking in Drying Colloidal Films. Phys. Rev. Lett. 2007, 98 (21), 218302, DOI: 10.1103/PhysRevLett.98.218302Google Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlvVagt78%253D&md5=ba2c452e115222337d1dee73f6c39e44Cracking in Drying Colloidal FilmsSingh, Karnail B.; Tirumkudulu, Mahesh S.Physical Review Letters (2007), 98 (21), 218302/1-218302/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)It has long been known that thick films of colloidal dispersions such as wet clays, paints, and coatings crack under drying. Although capillary stresses generated during drying have been recently identified as the cause for cracking, the existence of a max. crack-free film thickness that depends on particle size, rigidity, and packing has not been understood. Here, we identify two distinct regimes for crack-free films based on the magnitude of compressive strain at the max. attainable capillary pressure and show remarkable agreement of measurements with our theory. We anticipate our results to not only form the basis for design of coating formulations for the paints, coatings, and ceramics industry but also assist in the prodn. of crack-free photonic band gap crystals.
- 78Roberts, C. C.; Francis, L. F. Drying and Cracking of Soft Latex Coatings. J. Coatings Technol. Res. 2013, 10 (4), 441– 451, DOI: 10.1007/s11998-012-9425-7Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtVGrtL3I&md5=3a2d5ac7b9d2c3a16b6d2e00f51d7215Drying and cracking of soft latex coatingsRoberts, Christine C.; Francis, Lorraine F.Journal of Coatings Technology and Research (2013), 10 (4), 441-451CODEN: JCTRCP; ISSN:1935-3804. (Springer)The min. film formation temp. (MFFT) is the min. drying temp. needed for a latex coating to coalesce into an optically clear, dense crack-free film. To better understand the interplay of forces near this crit. temp., cryogenic SEM (cryoSEM) was used to track the latex particle deformation and water migration in coatings dried at temps. just above and below the MFFT. Although the latex particles completely coalesced at both temps. by the end of the drying process, it was discovered that particle deformation during the early drying stages was drastically different. Below the MFFT, cracks initiated just as menisci began to recede into the packing of consolidated particles, whereas above the MFFT, partial particle deformation occurred before menisci entered the coating and cracks were not obsd. The spacing between cracks measured in coatings dried at varying temps. decreased with decreasing drying temp. near the MFFT, whereas it was independent of temp. below a crit. temp. Finally, the addn. of small amts. of silica aggregates was found to lessen the cracking of latex coatings near the MFFT without adversely affecting their optical clarity.
- 79Cowie, J. M. G.; Arrighi, V. Polymers: Chemistry and Physics of Modern Materials, 3rd ed.; CRC Press: Boca Raton, 2007.Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 4 publications.
- Connor A. Sanders, Arthur Werner, Sandra E. Smeltzer, Sean R. George, Andreas Gernandt, Bernd Reck, Michael F. Cunningham. Amphiphilic Block Random Copolymers: Influence of pH and Ionic Strength on Aqueous Solution Properties. Macromolecules 2024, 57
(8)
, 3484-3495. https://doi.org/10.1021/acs.macromol.4c00095
- Devin Grabner, Phillip D. Pickett, Terry McAfee, Brian A. Collins. Molecular Weight-Independent “Polysoap” Nanostructure Characterized via In Situ Resonant Soft X-ray Scattering. Langmuir 2024, 40
(14)
, 7444-7455. https://doi.org/10.1021/acs.langmuir.3c03897
- Izabela Kurowska, Maksym Odnoroh, Oleksandr Ivanchenko, Marc Guerre, Mathias Destarac. Synthesis of high molar mass all-(meth)acrylic thermoplastic elastomers by photo-iniferter RAFT polymerisation. Polymer Chemistry 2024, 15
(47)
, 4888-4893. https://doi.org/10.1039/D4PY01123F
- Guoqing Jian, Ashok Santra, Christelle Mbuncha, Georgesha Ross. A novel star polymer for regulating fluid loss in oil-based mud under high temperature conditions. Journal of Molecular Liquids 2023, 389 , 122815. https://doi.org/10.1016/j.molliq.2023.122815
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 79 other publications.
- 1Bradford, E. B.; McKeever, L. D. Block Copolymers. Prog. Polym. Sci. 1971, 3, 109– 143, DOI: 10.1016/0079-6700(71)90003-71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE38XhtVSjtb8%253D&md5=c0d01b0a8f622d4c9873aa0e5f40137dBlock copolymersBradford, E. B.; McKeever, L. D.Progress in Polymer Science (1971), 3 (), 109-43CODEN: PRPSB8; ISSN:0079-6700.A review with 234 refs. discusses the synthesis, characterization, properties, and uses of block copolymers.
- 2Bates, F. S.; Fredrickson, G. H. Block Copolymers-Designer Soft Materials. Phys. Today 1999, 52 (2), 32– 38, DOI: 10.1063/1.8825222https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlymsr8%253D&md5=dd0e3bc303e671a3ae845969248fbc5eBlock copolymers-designer soft materialsBates, Frank S.; Fredrickson, Glenn H.Physics Today (1999), 52 (2), 32-38CODEN: PHTOAD; ISSN:0031-9228. (American Institute of Physics)A review with 18 refs. discusses the structure, morphol., phase behavior, and self-assembly properties of block copolymers.
- 3Amphiphilic Block Copolymers: Self-Assembly and Applications; Alexandridis, P.; Lindman, B., Eds.; Elsevier: Amsterdam, The Netherlands, 2000.There is no corresponding record for this reference.
- 4Zhang, L.; Yu, K.; Eisenberg, A. Ion-Induced Morphological Changes in “Crew-Cut” Aggregates of Amphiphilic Block Copolymers. Science 1996, 272, 1777– 1779, DOI: 10.1126/science.272.5269.17774https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjslWnurk%253D&md5=9c5f9e9eaeb98318d1d10c2a812316d3Ion-induced morphological changes in "crew-cut" aggregates of amphiphilic block copolymersZhang, Lifeng; Yu, Kui; Eisenberg, AdiScience (Washington, D. C.) (1996), 272 (5269), 1777-1779CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The addn. of ions in micromolar (CaCl2 or HCl) or millimolar (NaCl) concns. can change the morphol. of "crew-cut" aggregates of amphiphilic block copolymers in dil. solns. In addn. to spherical, rodlike, and univesicular or lamellar aggregates, an unusual large compd. vesicle morphol. can be obtained from a single block copolymer. Some features of the spontaneously formed large compd. vesicles may make them esp. useful as vehicles for delivering drugs and as models of biol. cells. Gelation of a dil. spherical micelle soln. can also be induced by ions as the result of the formation of a cross-linked "pearl necklace" morphol.
- 5Zhang, L.; Eisenberg, A. Multiple Morphologies of “Crew-Cut” Aggregates of Polystyrene-b-Poly(Acrylic Acid) Block Copolymers. Science 1995, 268, 1728– 1731, DOI: 10.1126/science.268.5218.17285https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmsFCisbk%253D&md5=dce426eb449e334f10438dbcf2d5aac6Multiple morphologies of "crew-cut" aggregates of polystyrene-b-poly(acrylic acid) block copolymersZhang, Lifeng; Eisenberg, AdiScience (Washington, D. C.) (1995), 268 (5218), 1728-31CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The observation by transmission electron microscopy of six different stable aggregate morphologies is reported for the same family of highly asym. styrene-acrylic acid diblock copolymers prepd. in a low-mol.-wt. solvent system. Four of the morphologies consist of spheres, rods, lamellae, and vesicles in aq. soln., whereas the fifth consists of simple reverse micelle-like aggregates. The sixth consists of up to micrometer-size spheres in aq. soln. that have hydrophilic surfaces and are filled with the reverse micelle-like aggregates. In addn., a needle-like solid, which is highly birefringent, is obtained on drying of aq. solns. of the spherical micelles. This range of morphologies is believed to be unprecedented for a block copolymer system.
- 6van Hest, J. C. M.; Delnoye, D. A. P.; Baars, M. W. P. L.; van Genderen, M. H. P.; Meijer, E. W. Polystyrene-Dendrimer Amphiphilic Block Copolymers with a Generation-Dependent Aggregation. Science 1995, 268 (268), 1592– 1595, DOI: 10.1126/science.268.5217.15926https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmtlarsbo%253D&md5=b4d4f8c6bbb3715ad64c022131a70bd7Polystyrene-dendrimer amphiphilic block copolymers with a generation-dependent aggregationvan Hest, J. C. M.; Delnoye, D. A. P.; Baars, M. W. P. L.; van Genderen, M. H. P.; Meijer, E. W.Science (Washington, D. C.) (1995), 268 (5217), 1592-5CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)A class of amphiphilic macromols. has been synthesized by combining well-defined polystyrene (PS) with poly(propylene imine) dendrimers. Five different generations, from PS-dendr-NH2 up to Ps-dendr-(NH2)32, were prepd. in yields of 70 to 90 percent. Dynamic light scattering, cond. measurements, and transmission electron microscopy show that in aq. phases, PS-dendr-(NH2)32 forms spherical micelles, PS-dendr-(NH2)16 forms micellar rods, and PS-dendr-(NH2)8 forms vesicular structures. The lower generations of this class of macromols. show inverted micellar behavior. The obsd. effect of amphiphile geometry on aggregation behavior is in qual. agreement with the theory of J. N. Israelachvili, et al. (1976). The amphiphiles presented here are similar in shape but different in size as compared with traditional surfactants, whereas they are similar in size but different in shape as compared with traditional block copolymers.
- 7Bates, F. S. Polymer-Polymer Phase Behaviour. Science 1991, 251 (251), 898– 905, DOI: 10.1126/science.251.4996.8987https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhsVCgsrk%253D&md5=53fc355f97cf74f3bf5e17e2c427509fPolymer-polymer phase behaviorBates, Frank S.Science (Washington, DC, United States) (1991), 251 (4996), 898-905CODEN: SCIEAS; ISSN:0036-8075.A review with 57 refs. on basic factors governing polymer-polymer phase behavior with emphasis on linear homopolymer blends and diblock copolymers as representative model mol. architectures. Equil. thermodn. and phase sepn. dynamics are reviewed and discussed.
- 8Mai, Y.; Eisenberg, A. Self-Assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41 (18), 5969– 5985, DOI: 10.1039/c2cs35115c8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1aqsbvL&md5=37964bf5011f9bc7a6aa42c4b612cf91Self-assembly of block copolymersMai, Yiyong; Eisenberg, AdiChemical Society Reviews (2012), 41 (18), 5969-5985CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Block copolymer (BCP) self-assembly has attracted considerable attention for many decades. The present tutorial review introduces the primary principles of BCP self-assembly in bulk and in soln., by describing expts., theories, accessible morphologies, etc.
- 9Xu, R.; Winnik, M. A.; Hallett, F. R.; Riess, G.; Croucher, M. D. Light-Scattering Study of the Association Behavior of Styrene-Ethylene Oxide Block Copolymers in Aqueous Solution. Macromolecules 1991, 24 (1), 87– 93, DOI: 10.1021/ma00001a0149https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXisFSmsQ%253D%253D&md5=1f8d2b01b4aafa5d0b14de8711b0384dLight-scattering study of the association behavior of styrene-ethylene oxide block copolymers in aqueous solutionXu, Renliang; Winnik, Mitchell A.; Hallett, F. R.; Riess, Gerard; Croucher, Melvin D.Macromolecules (1991), 24 (1), 87-93CODEN: MAMOBX; ISSN:0024-9297.The micellization behavior of styrene (I)-ethylene oxide (II) diblock and triblock copolymers in aq. solns. is studied via static and dynamic light scattering. There are 2 narrowly distributed populations of particles in the soln. The sizes of both sets of particles are unchanged in concn. range of 2 × 10 -5 - 2 × 10-3 g/mL. The wt. fraction of the larger particles decreases with increasing polymer concn. The small particles are regular micelles with an assocn. no. of several hundred; the large particles are loose micellar clusters of tens of micelles. The micelle structure is core-shell with a core of I and a shell of II.
- 10Wilhelm, M.; Zhao, C. Le; Wang, Y.; Xu, R.; Winnik, M. A.; Mura, J. L.; Riess, G.; Croucher, M. D. Poly(Styrene-Ethylene Oxide) Block Copolymer Micelle Formation in Water: A Fluorescence Probe Study. Macromolecules 1991, 24 (5), 1033– 1040, DOI: 10.1021/ma00005a01010https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhtVehtb8%253D&md5=5257329b56f4c306aa325112f726ae9fPoly(styrene-ethylene oxide) block copolymer micelle formation in water: a fluorescence probe studyWilhelm, Manfred; Zhao, Cheng Le; Wang, Yongcai; Xu, Renliang; Winnik, Mitchell A.; Mura, Jean Luc; Riess, Gerard; Croucher, Melvin D.Macromolecules (1991), 24 (5), 1033-40CODEN: MAMOBX; ISSN:0024-9297.Block copolymer micelle formation was studied by a combination of fluorescent and quasielastic light scattering (QELS) techniques. Ethylene oxide-styrene diblock and triblock copolymers, with mol. wts. 8500-29,000, formed spherical micelles in H2O over the entire concn. range over which QELS signals could be detected. Pyrene (I) in H2O (6 × 10-7 M) partitions between the aq. and micellar phases, accompanied by 3 changes in I spectroscopy. A red shift in the excitation spectrum, a change in the vibrational fine structure of I fluorescence, and an increase in the fluorescence decay time (from 200 to ∼350 ns) accompanied the transfer of I from an aq. to a hydrophobic micellar environment. From these data, crit. micelle concns. (1-5 mg/L) and partition coeffs. (3 × 105) were calcd.
- 11Prochazka, K.; Kiserow, D.; Ramireddy, C.; Tuzar, Z.; Munk, P.; Webber, S. E. Time-Resolved Fluorescence Studies of the Chain Dynamics of Naphthalene-Labeled Polystyrene-Block-Poly(Methacrylic Acid) Micelles in Aqueous Media. Macromolecules 1992, 25 (1), 454– 460, DOI: 10.1021/ma00027a07111https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38Xkt1aqug%253D%253D&md5=dc8198b84ce62c1791e10510da215520Time-resolved fluorescence studies of the chain dynamics of naphthalene-labeled polystyrene-block-poly(methacrylic acid) micelles in aqueous mediaProchazka, K.; Kiserow, D.; Ramireddy, C.; Tuzar, Z.; Munk, P.; Webber, S. E.Macromolecules (1992), 25 (1), 454-60CODEN: MAMOBX; ISSN:0024-9297.A-B block copolymers of polystyrene and poly(methacrylic acid) and polystyrene and poly(tert-Bu methacrylate), both with short-chain oligovinyl-2-naphthalene moieties attached to the end of the polystyrene block, were prepd. by anionic polymn. and micelles were formed by hydrolysis. Intramol. excimer formation (which is controlled by the mobility of the pendant fluorescent groups and the polymer chain dynamics) was sterically hindered in the micellar cores as they become more compact. Both steady-state and time-resolved excimer fluorescence were significantly influenced by the gradual collapse of the micellar cores and the increase in segment d. within the cores with an increasing content of water. Changes in lifetimes and preexponential factors for naphthalene fluorescence (monomer as well as excimer) were sensitive indicators of micelle formation.
- 12Astafieva, I.; Zhong, X. F.; Eisenberg, A. Critical Micellization Phenomena in Block Polyelectrolyte Solutions. Macromolecules 1993, 26 (26), 7339– 7352, DOI: 10.1021/ma00078a03412https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmslahur4%253D&md5=bd163a0318c8c488cb225d305f2ffc50Critical micellization phenomena in block polyelectrolyte solutionsAstafieva, Irina; Zhong, Xing Fu; Eisenberg, AdiMacromolecules (1993), 26 (26), 7339-52CODEN: MAMOBX; ISSN:0024-9297.Crit. micelle concn. (CMC) was measured for a range of block copolyelectrolytes based on styrene (the insol. block) and Na acrylate. The length of the styrene blocks was 6-110, while that of the polyelectrolyte was ∼300-1400. The fluorescent probe technique was used with pyrene as a probe mol., and the data were treated by 5 different methods. CMC results were interpolated for a const. polyelectrolyte block length of 1000 units. Changing the insol. block length from 6 to 110 lowed the CMC from 1.6 × 10-5 to 5 × 10-8 M. By contrast, changing the sol. block length from 300 to 1400 typically changed the CMC values by a factor of <2. For very short polystyrene block lengths, the CMC decreased very rapidly with increasing length of the insol. block. By contrast, for higher block lengths (>12) the drop in the CMC was much more gradual. The results are compared with those of previous studies of block copolymer micelles and theories of block copolymer micellization.
- 13Qin, A.; Tian, M.; Ramireddy, C.; Webber, S. E.; Munk, P.; Tuzar, Z. Polystyrene-Poly(Methacrylic Acid) Block Copolymer Micelles. Macromolecules 1994, 27 (1), 120– 126, DOI: 10.1021/ma00079a01813https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXjtFOltw%253D%253D&md5=38e7c164b285a85c5e089d0da6b234aaPolystyrene-poly(methacrylic acid) block copolymer micellesQin, Anwei; Tian, Minmin; Ramireddy, C.; Webber, Stephen E.; Munk, Petr; Tuzar, ZdenekMacromolecules (1994), 27 (1), 120-6CODEN: MAMOBX; ISSN:0024-9297.Block copolymer micelles formed by diblock and triblock copolymers of styrene (I) and methacrylic acid (II) were characterized in soln. in a mixed solvent with 80 vol % of dioxane and 20 vol % of water. Methods of static light scattering, quasielastic light scattering, differential refractometry, viscometry, sedimentation velocity, and densitometry were used. No unattached unimer mols. were obsd. Three independent methods were employed for obtaining micellar wts. and agreed well with each other. No anomalous behavior was obsd. by any method. The micellar solns. contained almost exclusively single micelles; only a few samples (those producing the largest micelles) contained micellar clusters. The micelles behaved hydrodynamically and thermodynamically as impermeable spheres. The structure of the micellar shell was characterized in some detail. Relations between the aggregation no. and the hydrodynamic radius of the micelles on the one hand and the sizes of the I and II blocks on the other were presented as scaling-type phenomenol. equations.
- 14Guo, M.; Pitet, L. M.; Wyss, H. M.; Vos, M.; Dankers, P. Y. W.; Meijer, E. W. Tough Stimuli-Responsive Supramolecular Hydrogels with Hydrogen-Bonding Network Junctions. J. Am. Chem. Soc. 2014, 136 (19), 6969– 6977, DOI: 10.1021/ja500205v14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXns1CntLY%253D&md5=ad1f36275bc00ec6050c57e2e6973716Tough Stimuli-Responsive Supramolecular Hydrogels with Hydrogen-Bonding Network JunctionsGuo, Mingyu; Pitet, Louis M.; Wyss, Hans M.; Vos, Matthijn; Dankers, Patricia Y. W.; Meijer, E. W.Journal of the American Chemical Society (2014), 136 (19), 6969-6977CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Hydrogels were prepd. with phys. crosslinks comprising 2-ureido-4[1H]-pyrimidinone (UPy) hydrogen-bonding units within the backbone of segmented amphiphilic macromols. contg. hydrophilic poly(ethylene glycol) (PEG). The bulk materials adopt nanoscopic phys. crosslinks composed of UPy-UPy dimers embedded in segregated hydrophobic domains dispersed within the PEG matrix as confirmed by cryo-electron microscopy. The amphiphilic network was swollen with high wt. fractions of water (wH2O ≈ 0.8) owing to the high PEG wt. fraction within the pristine polymers (wPEG ≈ 0.9). Two different PEG chain lengths were investigated and illustrate the corresponding consequences of crosslink d. on mech. properties. The resulting hydrogels exhibited high strength and resilience upon deformation, consistent with a microphase sepd. network, in which the UPy-UPy interactions were adequately shielded within hydrophobic nanoscale pockets that maintain the network despite extensive water content. The cumulative result is a series of tough hydrogels with tunable mech. properties and tractable synthetic prepn. and processing. Furthermore, the melting transition of PEG in the dry polymer was shown to be an effective stimulus for shape memory behavior.
- 15Huo, M.; Yuan, J.; Tao, L.; Wei, Y. Redox-Responsive Polymers for Drug Delivery: From Molecular Design to Applications. Polym. Chem. 2014, 5 (5), 1519– 1528, DOI: 10.1039/C3PY01192E15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvFGhs74%253D&md5=87de58de0d857d687802195ecafd49c6Redox-responsive polymers for drug delivery: from molecular design to applicationsHuo, Meng; Yuan, Jinying; Tao, Lei; Wei, YenPolymer Chemistry (2014), 5 (5), 1519-1528CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A review. Glutathione has been regarded as a significant signal for distinguishing between tumor and normal tissue. Recently, reactive oxygen species have attracted much attention for their close connection with many diseases. Taking advantage of the physiol. signals, redox-responsive polymeric drug carriers constitute a significant research area in the various stimuli-responsive polymers for biomedical applications. During the rapid development of redox-responsive polymers, mol. design and related synthetic methodol. plays a crucial role. In this review, we discuss the redn.- and oxidn.-responsive polymeric drug carriers from the view of functional groups, as well as their applications in controlled release.
- 16Schattling, P.; Jochum, F. D.; Theato, P. Multi-Stimuli Responsive Polymers - the All-in-One Talents. Polym. Chem. 2014, 5 (1), 25– 36, DOI: 10.1039/C3PY00880K16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvVKjtLfJ&md5=9de59f5913441e9f2b3e7c77d803bb85Multi-stimuli responsive polymers - the all-in-one talentsSchattling, Philipp; Jochum, Florian D.; Theato, PatrickPolymer Chemistry (2014), 5 (1), 25-36CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Stimuli-responsive polymers have gained increasing attention, which is attributed to the manifold applications they can be used for. Several years' intensive research was invested in stimuli-responsive polymers. Their stimuli-responsiveness led not only to novel responsive groups, which enabled the translation of an external phys. impact into a change of a material property, but also to polymers that are equipped with more than one responsive group. The integration of several responsive moieties within one polymer yields smart polymers exhibiting complex responsive behavior of the polymers. This review summarises recent developments in the area of multi-stimuli responsive polymers, laying the focus on the improved, multifaceted response of polymer materials depending on the impact of several external stimuli.
- 17Sun, Y.; Wang, Z.; Li, Y.; Zhang, Z.; Zhang, W.; Pan, X.; Zhou, N.; Zhu, X. Photoresponsive Amphiphilic Macrocycles Containing Main-Chain Azobenzene Polymers. Macromol. Rapid Commun. 2015, 36, 1341– 1347, DOI: 10.1002/marc.20150013617https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXnvFWmtr4%253D&md5=90de703b2124052f0c097b06a8bd09a0Photoresponsive Amphiphilic Macrocycles Containing Main-Chain Azobenzene PolymersSun, Yadong; Wang, Zhao; Li, Yiwen; Zhang, Zhengbiao; Zhang, Wei; Pan, Xiangqiang; Zhou, Nianchen; Zhu, XiulinMacromolecular Rapid Communications (2015), 36 (14), 1341-1347CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, the first example of photosensitive cyclic amphiphilic homopolymers consisting of multiple biphenyl azobenzene chromophores in the cyclic main chain tethered with hydrophilic tetraethylene glycol monomethyl ether units is presented. The synthetic approach involves sequentially performed thermal catalyzed "click" step-growth polymn. in bulk, and Cu(I)-catalyzed azide-alkyne cycloaddn. (CuAAC) intramol. cyclization from α-alkyne/ω-azide linear precursors. It is obsd. that such amphiphilic macrocycles exhibit increased glass transition temps. (Tg), slightly faster trans-cis-trans photoisomerization, and enhanced fluorescence emission intensity compared with the corresponding linear polymers. In addn., the cyclic amphiphilic homopolymers self-assemble into spherical nanoparticles with smaller sizes which possess slower photoresponsive behaviors in a tetrahydrofuran/water mixt. compared with those of the linear ones. All these interesting observations suggest that the cyclic topol. has a great influence on the phys. properties and self-assembly behavior of these photoresponsive amphiphilic macrocycles in general.
- 18Blanazs, A.; Madsen, J.; Battaglia, G.; Ryan, A. J.; Armes, S. P. Mechanistic Insights for Block Copolymer Morphologies: How Do Worms Form Vesicles?. J. Am. Chem. Soc. 2011, 133, 16581– 16587, DOI: 10.1021/ja206301a18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFCgsLfM&md5=2f1761acb3dac1734f68866820e00119Mechanistic Insights for Block Copolymer Morphologies: How Do Worms Form Vesicles?Blanazs, Adam; Madsen, Jeppe; Battaglia, Giuseppe; Ryan, Anthony J.; Armes, Steven P.Journal of the American Chemical Society (2011), 133 (41), 16581-16587CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Amphiphilic diblock copolymers composed of two covalently linked, chem. distinct chains can be considered to be biol. mimics of cell membrane-forming lipid mols., but with typically more than an order of magnitude increase in mol. wt. These macromol. amphiphiles are known to form a wide range of nanostructures (spheres, worms, vesicles, etc.) in solvents that are selective for one of the blocks. However, such self-assembly is usually limited to dil. copolymer solns. (<1%), which is a significant disadvantage for potential com. applications such as drug delivery and coatings. In principle, this problem can be circumvented by polymn.-induced block copolymer self-assembly. Here the authors detail the synthesis and subsequent in situ self-assembly of amphiphilic AB diblock copolymers in a one pot concd. aq. dispersion polymn. formulation. The authors show that spherical micelles, wormlike micelles, and vesicles can be predictably and efficiently obtained (within 2 h of polymn., >99% monomer conversion) at relatively high solids in purely aq. soln. Furthermore, careful monitoring of the in situ polymn. by transmission electron microscopy reveals various novel intermediate structures (including branched worms, partially coalesced worms, nascent bilayers, "octopi", "jellyfish", and finally pure vesicles) that provide important mechanistic insights regarding the evolution of the particle morphol. during the sphere-to-worm and worm-to-vesicle transitions. This environmentally benign approach (which involves no toxic solvents, is conducted at relatively high solids, and requires no addnl. processing) is readily amenable to industrial scale-up, since it is based on com. available starting materials.
- 19Lovett, J. R.; Ratcliffe, L. P. D.; Warren, N. J.; Armes, S. P.; Smallridge, M. J.; Cracknell, R. B.; Saunders, B. R. A Robust Cross-Linking Strategy for Block Copolymer Worms Prepared via Polymerization-Induced Self-Assembly. Macromolecules 2016, 49 (8), 2928– 2941, DOI: 10.1021/acs.macromol.6b0042219https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XmtVWisL0%253D&md5=6c72e71bec9902b7aab48eb278fcdb53A Robust Cross-Linking Strategy for Block Copolymer Worms Prepared via Polymerization-Induced Self-AssemblyLovett, J. R.; Ratcliffe, L. P. D.; Warren, N. J.; Armes, S. P.; Smallridge, M. J.; Cracknell, R. B.; Saunders, B. R.Macromolecules (Washington, DC, United States) (2016), 49 (8), 2928-2941CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A poly(glycerol monomethacrylate) (PGMA) chain transfer agent is chain-extended by reversible addn.-fragmentation chain transfer (RAFT) statistical copolymn. of 2-hydroxypropyl methacrylate (HPMA) with glycidyl methacrylate (GlyMA) in concd. aq. soln. via polymn.-induced self-assembly (PISA). A series of five free-standing worm gels is prepd. by fixing the overall d.p. of the core-forming block at 144 while varying its GlyMA content from 0 to 20 mol %. 1H NMR kinetics indicated that GlyMA is consumed much faster than HPMA, producing a GlyMA-rich sequence close to the PGMA stabilizer block. Temp.-dependent oscillatory rheol. studies indicate that increasing the GlyMA content leads to progressively less thermoresponsive worm gels, with no degelation on cooling being obsd. for worms contg. 20 mol % GlyMA. The epoxy groups in the GlyMA residues can be ring-opened using 3-aminopropyltriethoxysilane (APTES) in order to prep. core crosslinked worms via hydrolysis-condensation with the siloxane groups and/or hydroxyl groups on the HPMA residues. Perhaps surprisingly, 1H NMR anal. indicates that the epoxy-amine reaction and the intermol. crosslinking occur on similar time scales. Crosslinking leads to stiffer worm gels that do not undergo degelation upon cooling. Dynamic light scattering studies and TEM analyses conducted on linear worms exposed to either methanol (a good solvent for both blocks) or anionic surfactant result in immediate worm dissocn. In contrast, crosslinked worms remain intact under such conditions, provided that the worm cores comprise at least 10 mol % GlyMA.
- 20Blanazs, A.; Ryan, A. J.; Armes, S. P. Predictive Phase Diagrams for RAFT Aqueous Dispersion Polymerization: Effect of Block Copolymer Composition, Molecular Weight, and Copolymer Concentration. Macromolecules 2012, 45 (12), 5099– 5107, DOI: 10.1021/ma301059r20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XosFemtrc%253D&md5=5847f249ec070800f649a2a2f13a2b86Predictive Phase Diagrams for RAFT Aqueous Dispersion Polymerization: Effect of Block Copolymer Composition, Molecular Weight, and Copolymer ConcentrationBlanazs, A.; Ryan, A. J.; Armes, S. P.Macromolecules (Washington, DC, United States) (2012), 45 (12), 5099-5107CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Polymn.-induced self-assembly (PISA) of poly(glycerol monomethacrylate)-poly(2-hydroxypropyl methacrylate) (PGMA-PHPMA) diblocks is conducted using a RAFT aq. dispersion polymn. formulation at 70 °C. Several PGMA macromol. chain transfer agents (macro-CTAs) are chain-extended using a water-miscible monomer (HPMA): the growing PHPMA block becomes increasingly hydrophobic and hence drives in situ self-assembly. The final copolymer morphol. in such PISA syntheses depends on just three parameters: the mean d.p. (DP) of the PGMA stabilizer block, the mean DP of the PHPMA core-forming block, and the total solids concn. TEM is used to construct detailed diblock copolymer phase diagrams for PGMA DPs of 47, 78, and 112. For the shortest stabilizer block, there is essentially no concn. dependence: spheres, worms, or vesicles can be obtained even at 10% wt./wt. solids simply by selecting the DP of the PHPMA block that gives the appropriate mol. curvature. For a PGMA DP of 78, the phase diagram is rich: and the copolymer morphol. depends strongly on the total solids concn. There is also a narrow region where spheres, worms, and vesicles coexist, which may be due to the effect of polydispersity. For a PGMA112 macro-CTA, the phase diagram is dominated by spherical morphologies. This is probably because the longer core-forming block DPs required to reduce the mol. curvature are significantly more dehydrated and hence less mobile, which prevents the in situ evolution of morphol. from spheres to higher order morphologies. This hypothesis is supported by the observation that addn. of ethanol to aq. PISA syntheses conducted using the longer macro-CTAs allows access to diblock copolymer worms or vesicles, since this cosolvent solvates the core-forming PHPMA chains and hence increases their mobility at 70 °C. Elucidation of such phase diagrams is vital to ensure reproducible targeting of pure phases, rather than mixed phases.
- 21Mable, C. J.; Gibson, R. R.; Prevost, S.; McKenzie, B. E.; Mykhaylyk, O. O.; Armes, S. P. Loading of Silica Nanoparticles in Block Copolymer Vesicles during Polymerization-Induced Self-Assembly: Encapsulation Efficiency and Thermally Triggered Release. J. Am. Chem. Soc. 2015, 137 (51), 16098– 16108, DOI: 10.1021/jacs.5b1041521https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFSmtb%252FK&md5=3cc548f55bc0f03245e24924f8657210Loading of Silica Nanoparticles in Block Copolymer Vesicles during Polymerization-Induced Self-Assembly: Encapsulation Efficiency and Thermally Triggered ReleaseMable, Charlotte J.; Gibson, Rebecca R.; Prevost, Sylvain; McKenzie, Beulah E.; Mykhaylyk, Oleksandr O.; Armes, Steven P.Journal of the American Chemical Society (2015), 137 (51), 16098-16108CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Poly(glycerol monomethacrylate)-poly(2-hydroxypropyl methacrylate) diblock copolymer vesicles can be prepd. in the form of concd. aq. dispersions via polymn.-induced self-assembly (PISA). In the present study, these syntheses are conducted in the presence of varying amts. of silica nanoparticles of approx. 18 nm diam. This approach leads to encapsulation of up to hundreds of silica nanoparticles per vesicle. Silica has high electron contrast compared to the copolymer which facilitates TEM anal., and its thermal stability enables quantification of the loading efficiency via thermogravimetric anal. Encapsulation efficiencies can be calcd. using disk centrifuge photosedimentometry, since the vesicle d. increases at higher silica loadings while the mean vesicle diam. remains essentially unchanged. Small angle X-ray scattering (SAXS) is used to confirm silica encapsulation, since a structure factor is obsd. at q ≈ 0.25 nm-1. A new two-population model provides satisfactory data fits to the SAXS patterns and allows the mean silica vol. fraction within the vesicles to be detd. Finally, the thermoresponsive nature of the diblock copolymer vesicles enables thermally triggered release of the encapsulated silica nanoparticles simply by cooling to 0-10 °C, which induces a morphol. transition. These silica-loaded vesicles constitute a useful model system for understanding the encapsulation of globular proteins, enzymes, or antibodies for potential biomedical applications. They may also serve as an active payload for self-healing hydrogels or repair of biol. tissue. Finally, we also encapsulate a model globular protein, bovine serum albumin, and calc. its loading efficiency using fluorescence spectroscopy.
- 22Rösler, A.; Vandermeulen, G. W. M.; Klok, H. A. Advanced Drug Delivery Devices via Self-Assembly of Amphiphilic Block Copolymers. Adv. Drug Delivery Rev. 2012, 64, 270– 279, DOI: 10.1016/j.addr.2012.09.026There is no corresponding record for this reference.
- 23Derry, M. J.; Mykhaylyk, O. O.; Armes, S. P. A Vesicle-to-Worm Transition Provides a New High-Temperature Oil Thickening Mechanism. Angew. Chemie - Int. Ed. 2017, 56 (7), 1746– 1750, DOI: 10.1002/anie.20160936523https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlsFCjtg%253D%253D&md5=c21c3e5268d7378a9d75321f947e0589A Vesicle-to-Worm Transition Provides a New High-Temperature Oil Thickening MechanismDerry, Matthew J.; Mykhaylyk, Oleksandr O.; Armes, Steven P.Angewandte Chemie, International Edition (2017), 56 (7), 1746-1750CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Diblock copolymer vesicles are prepd. via RAFT dispersion polymn. directly in mineral oil. Such vesicles undergo a vesicle-to-worm transition on heating to 150 °C, as judged by TEM and SAXS. Variable-temp. 1H NMR spectroscopy indicates that this transition is the result of surface plasticization of the membrane-forming block by hot solvent, effectively increasing the vol. fraction of the stabilizer block and so reducing the packing parameter for the copolymer chains. The rheol. behavior of a 10 % wt./wt. copolymer dispersion in mineral oil is strongly temp.-dependent: the storage modulus increases by five orders of magnitude on heating above the crit. gelation temp. of 135 °C, as the non-interacting vesicles are converted into weakly interacting worms. SAXS studies indicate that, on av., three worms are formed per vesicle. Such vesicle-to-worm transitions offer an interesting new mechanism for the high-temp. thickening of oils.
- 24Bates, F. S. Polymer-Polymer Phase Behavior. Science (80-.). 1991, 251 (4996), 898– 905, DOI: 10.1126/science.251.4996.89824https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhsVCgsrk%253D&md5=53fc355f97cf74f3bf5e17e2c427509fPolymer-polymer phase behaviorBates, Frank S.Science (Washington, DC, United States) (1991), 251 (4996), 898-905CODEN: SCIEAS; ISSN:0036-8075.A review with 57 refs. on basic factors governing polymer-polymer phase behavior with emphasis on linear homopolymer blends and diblock copolymers as representative model mol. architectures. Equil. thermodn. and phase sepn. dynamics are reviewed and discussed.
- 25Blanazs, A.; Armes, S. P.; Ryan, A. J. Self-Assembled Block Copolymer Aggregates: From Micelles to Vesicles and Their Biological Applications. Macromol. Rapid Commun. 2009, 30 (4), 267– 277, DOI: 10.1002/marc.20080071325https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXivFelu7k%253D&md5=a7578981e5867fcee9d50b9e361c9597Self-assembled block copolymer aggregates: from micelles to vesicles and their biological applicationsBlanazs, Adam; Armes, Steven P.; Ryan, Anthony J.Macromolecular Rapid Communications (2009), 30 (4-5), 267-277CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The ability of amphiphilic block copolymers to self-assemble in selective solvents has been widely studied in academia and utilized for various com. products. The self-assembled polymer vesicle is at the forefront of this nanotechnol. revolution with seemingly endless possible uses, ranging from biomedical to nanometer-scale enzymic reactors. This review is focused on the inherent advantages in using polymer vesicles over their small mol. lipid counterparts and the potential applications in biol. for both drug delivery and synthetic cellular reactors.
- 26Albigès, R.; Klein, P.; Roi, S.; Stoffelbach, F.; Creton, C.; Bouteiller, L.; Rieger, J. Water-Based Acrylic Coatings Reinforced by PISA-Derived Fibers. Polym. Chem. 2017, 8 (34), 4992– 4995, DOI: 10.1039/C7PY00302A26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmtFGnt70%253D&md5=7c7acccbb5661d2aad39d9b2b74729d3Water-based acrylic coatings reinforced by PISA-derived fibersAlbiges, Richard; Klein, Pauline; Roi, Stephanie; Stoffelbach, Francois; Creton, Costantino; Bouteiller, Laurent; Rieger, JuttaPolymer Chemistry (2017), 8 (34), 4992-4995CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)Polymn.-induced self-assembly (PISA) is an efficient approach to synthesize self-assembled core-shell nanofibers in water, made of amphiphilic block copolymers. We demonstrate here for the first time that such anisotropic and high Tg nanofibers can be used as reinforcing fillers for water-based acrylic materials. This strategy opens the way for strong all-org. aq. coatings.
- 27Chenal, M.; Rieger, J.; Vechambre, C.; Chenal, J.-M.; Chazeau, L.; Creton, C.; Bouteiller, L. Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard Phase. Macromol. Rapid Commun. 2013, 34 (34), 1524– 1529, DOI: 10.1002/marc.20130041027https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlahtLbO&md5=4d4bc64299888278198f8cdf39b17d57Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard PhaseChenal, Marion; Rieger, Jutta; Vechambre, Cyril; Chenal, Jean-Marc; Chazeau, Laurent; Creton, Costantino; Bouteiller, LaurentMacromolecular Rapid Communications (2013), 34 (19), 1524-1529CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)In this study, aq. emulsion polymn. of Bu acrylate is performed in batch conditions without surfactants using a poly(acrylic acid)-trithiocarbonate macro-RAFT agent to control the polymn. and to stabilize the emulsion. According to the polymn.-induced self-assembly (PISA) approach, well-defined amphiphilic PAA-b-PBA diblock copolymers form and self-assemble during synthesis to yield highly stable core-shell particles with an extremely thin hard PAA shell. For the first time, we report here the specific properties of films obtained from these particular latexes. After drying the aq. dispersion, tough and transparent films are obtained. Although the films are not chem. crosslinked, they do not dissolve in good solvents for PBA. Moreover, they remain transparent even after immersion in water. Rheol. shows that the films are both stiff and ductile, thanks to the nanostructured but very low vol. fraction (less than 3 wt%) of PAA forming a percolating network in the soft PBA. Compared with conventional core-shell-based films, this approach affords for the first time a route to a thin percolating honeycomb nanostructure with a sharp and strong interface between the two phases. The versatility of the synthetic procedure opens perspectives for a large range of functional materials.
- 28Rieger, J.; Osterwinter, G.; Bui, C.; Stoffelbach, F.; Charleux, B. Surfactant-Free Controlled/Living Radical Emulsion (Co)Polymerization of n-Butyl Acrylate and Methyl Methacrylate via RAFT Using Amphiphilic Polyethylene Oxide-Based Trithiocarbonate Chain Transfer Agents. Macromolecules 2009, 42 (15), 5518– 5525, DOI: 10.1021/ma900880328https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXotVGlt74%253D&md5=d13143067792442f9941957db3ccc8d9Surfactant-Free Controlled/Living Radical Emulsion (Co)polymerization of n-Butyl Acrylate and Methyl Methacrylate via RAFT Using Amphiphilic Poly(ethylene oxide)-Based Trithiocarbonate Chain Transfer AgentsRieger, Jutta; Osterwinter, Gregor; Bui, Chuong; Stoffelbach, Francois; Charleux, BernadetteMacromolecules (Washington, DC, United States) (2009), 42 (15), 5518-5525CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The RAFT-mediated, surfactant-free, ab initio, batch emulsion polymn. of Bu acrylate (nBA) and its copolymn. with Me methacrylate (MMA) were studied. The control agent was a surface-active trithiocarbonate macromol. RAFT agent composed of a hydrophilic poly(ethylene oxide) (PEO) block and a hydrophobic dodecyl chain. The homopolymns. of nBA were fast with high final conversions, and the polymer chains were well-controlled with narrow molar mass distribution. The length of the PEO chain was shown to affect the particle size and the polymn. kinetics directly. We found the conditions to tune the particle size independently from the poly(Bu acrylate) chain length by playing with a mixt. of macro-RAFT agents with long and short PEO segment or by adding a PEO-based nonionic surfactant. The copolymns. of nBA and MMA exhibited features very similar to those of the nBA homopolymns. provided that the molar percentage of MMA did not exceed approx. 75%. In all cases, stable, submicrometric particles composed of amphiphilic diblock copolymer chains were formed.
- 29Cockram, A. A.; Bradley, R. D.; Lynch, S. A.; Fleming, P. C. D.; Williams, N. S. J.; Murray, M. W.; Emmett, S. N.; Armes, S. P. Optimization of the High-Throughput Synthesis of Multiblock Copolymer Nanoparticles in Aqueous Media: Via Polymerization-Induced Self-Assembly. React. Chem. Eng. 2018, 3 (5), 645– 657, DOI: 10.1039/C8RE00066B29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtFCjsrfF&md5=b4951aaeaec274ad421b06f6c01670afOptimization of the high-throughput synthesis of multiblock copolymer nanoparticles in aqueous media via polymerization-induced self-assemblyCockram, Amy A.; Bradley, Robert D.; Lynch, Sylvie A.; Fleming, Patricia C. D.; Williams, Neal S. J.; Murray, Martin W.; Emmett, Simon N.; Armes, Steven P.Reaction Chemistry & Engineering (2018), 3 (5), 645-657CODEN: RCEEBW; ISSN:2058-9883. (Royal Society of Chemistry)In the present study, we report that PISA formulations are sufficiently robust to enable high-throughput expts. using a com. synthesis robot (Chemspeed Autoplant A100). More specifically, we use reversible addn.-fragmentation chain transfer (RAFT) aq. emulsion polymn. of either Bu methacrylate and/or benzyl methacrylate to prep. various examples of methacrylic multiblock copolymer nanoparticles using a poly(methacrylic acid) stabilizer block. Adequate stirring is essential to generate sufficiently small monomer droplets for such heterogeneous polymns. to proceed efficiently. Good reproducibility can be achieved under such conditions, with well-defined spherical morphologies being obtained at up to 45% wt./wt. solids. GPC studies indicate high blocking efficiencies but relatively broad mol. wt. distributions (Mw/Mn = 1.36-1.85), suggesting well-defined (albeit rather polydisperse) block copolymer chains. These preliminary studies provide a sound basis for high-throughput screening of RAFT-mediated PISA formulations, which is likely to be required for commercialization of this technol. Our results indicate that methacrylic PISA formulations enable the synthesis of diblock and triblock copolymer nanoparticles in high overall yield (94-99%) within 1-3 h at 70 °C. However, tetrablocks suffer from incomplete conversions (87-96% within 5 h) and hence most likely represent the upper limit for this approach.
- 30Cockram, A. A.; Neal, T. J.; Derry, M. J.; Mykhaylyk, O. O.; Williams, N. S. J.; Murray, M. W.; Emmett, S. N.; Armes, S. P. Effect of Monomer Solubility on the Evolution of Copolymer Morphology during Polymerization-Induced Self-Assembly in Aqueous Solution. Macromolecules 2017, 50 (3), 796– 802, DOI: 10.1021/acs.macromol.6b0230930https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsFKms7w%253D&md5=cd3a26b52453e4fa9ab4f727e5b8a9feEffect of Monomer Solubility on the Evolution of Copolymer Morphology during Polymerization-Induced Self-Assembly in Aqueous SolutionCockram, Amy A.; Neal, Thomas J.; Derry, Matthew J.; Mykhaylyk, Oleksandr O.; Williams, Neal S. J.; Murray, Martin W.; Emmett, Simon N.; Armes, Steven P.Macromolecules (Washington, DC, United States) (2017), 50 (3), 796-802CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Polymn.-induced self-assembly (PISA) has become a widely used technique for the design of diblock copolymer nano-objects in concd. aq. soln. Depending on the specific PISA formulation, reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. typically provides straightforward access to either spheres, worms or vesicles. In contrast, RAFT aq. emulsion polymn. formulations often lead to just kinetically trapped spheres. This limitation is currently not understood and only a few empirical exceptions were reported in the literature. In the present work, the effect of monomer soly. on copolymer morphol. is explored for an aq. polymn.-induced self-assembly (PISA) formulation. Using 2-hydroxybutyl methacrylate (aq. soly. =20 g dm-3 at 70 °C) instead of benzyl methacrylate (0.40 g dm-3 at 70 °C) for the core-forming block allows access to an unusual "monkey nut" copolymer morphol. over a relatively narrow range of target ds.p. when using a poly(methacrylic acid) RAFT agent at pH 5. These new nanoparticles have been characterized by transmission electron microscopy, dynamic light scattering, aq. electrophoresis, shear-induced polarized light imaging (SIPLI) and small-angle X-ray scattering.
- 31Rieger, J. Guidelines for the Synthesis of Block Copolymer Particles of Various Morphologies by RAFT Dispersion Polymerization. Macromol. Rapid Commun. 2015, 36 (16), 1458– 1471, DOI: 10.1002/marc.20150002831https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXovVCku78%253D&md5=bef40a4da086930c94c52099f796b0bcGuidelines for the Synthesis of Block Copolymer Particles of Various Morphologies by RAFT Dispersion PolymerizationRieger, JuttaMacromolecular Rapid Communications (2015), 36 (16), 1458-1471CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This article presents the recent developments of radical dispersion polymerizaton controlled by reversible addn. fragmentation chain transfer (RAFT) for the prodn. of block copoly-mer particles of various morphologies, such as core-shell spheres, worms, or vesicles. It is not meant to be an exhaustive review but it rather provides guidelines for non-specialists. The article is subdivided into eight sections. After a general introduction, the mechanism of polymn.-induced self-assembly (PISA) through RAFT-mediated dispersion polymn. is presented and the different parameters that control the morphol. produced are discussed. The next two sections are devoted to the choice of the monomer/solvent pair and the macroRAFT agent. Afterwards, post-polymn. morphol. order-to-order transitions (i.e. morphol. transitions triggered by extrinsic stimuli) or order-to-disorder transitions (i.e. disassembly of chains) are discussed. Assemblies based on more complex polymer architectures, such as triblock copolymers, are presented next, and finally the possibility to stabilize these structures by crosslinking is reported. The manuscript ends with a short conclusion and an outlook.
- 32Penfold, N. J. W.; Whatley, J. R.; Armes, S. P. Thermoreversible Block Copolymer Worm Gels Using Binary Mixtures of PEG Stabilizer Blocks. Macromolecules 2019, 52 (4), 1653– 1662, DOI: 10.1021/acs.macromol.8b0249132https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXis1antbw%253D&md5=0f5272986cdd505e3a655d03c7d37140Thermoreversible Block Copolymer Worm Gels Using Binary Mixtures of PEG Stabilizer BlocksPenfold, Nicholas J. W.; Whatley, Jessica R.; Armes, Steven P.Macromolecules (Washington, DC, United States) (2019), 52 (4), 1653-1662CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Two trithiocarbonate-based poly(ethylene glycol) (PEG) macromol. chain transfer agents (macro-CTAs) with mean ds.p. of 45 and 113 were prepd. with ≥94% chain-end functionality. Binary mixts. of these PEG-trithiocarbonate macro-CTAs were then chain-extended via reversible addn.-fragmentation chain transfer (RAFT) aq. dispersion polymn. of 2-hydroxypropyl methacrylate (HPMA). Systematic variation of the relative proportions of PEG45 and PEG113 macro-CTAs and the d.p. of the PHPMA core-forming block resulted in the formation of [x PEG45 + z PEG113] - PHPMAn block copolymer spheres, worms, or vesicles, where x and z represent the mole fractions of PEG45 and PEG113, resp. A phase diagram was constructed to establish the relationship between block copolymer compn. and nanoparticle morphol. The thermoresponsive behavior of block copolymer worms was assessed by visual inspection, dynamic light scattering (DLS), transmission electron microscopy (TEM) and temp.-dependent oscillatory rheol. Increasing the proportion of PEG45 (x = 0.00-0.40) in the stabilizer block resulted in a moderate increase in worm gel strength, but cooling resulted in irreversible degelation owing to a worm-to-sphere morphol. transition. However, the phase diagram enabled identification of a single diblock copolymer compn. that exhibited reversible degelation behavior in pure water. This formulation was then further optimized to exhibit the same rheol. behavior in a com. cell culture medium (Nutristem) by fixing the PEG mole fraction at x = 0.70 while lowering the PHPMA DP from 115 to 75. Importantly, the gel strength at physiol. temp. can be readily tuned simply by variation of the copolymer concn. In principle, this study has important implications for the preservation of human stem cells, which can enter stasis when immersed in certain worm gels [see: Canton et al. ACS Cent. Sci.2016, 2, 65-74].
- 33Qiao, Z.; Qiu, T.; Liu, W.; Zhang, L.; Tu, J.; Guo, L.; Li, X. A “Green” Method for Preparing ABCBA Penta-Block Elastomers by Using RAFT Emulsion Polymerization. Polym. Chem. 2017, 8 (19), 3013– 3021, DOI: 10.1039/C7PY00464H33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXmtVSis7w%253D&md5=2bfb612e9ce191da9a20f7d2843f94feA "green" method for preparing ABCBA penta-block elastomers by using RAFT emulsion polymerizationQiao, Zhi; Qiu, Teng; Liu, Weiwei; Zhang, Liangdong; Tu, Jinqiang; Guo, Longhai; Li, XiaoyuPolymer Chemistry (2017), 8 (19), 3013-3021CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)A clean method for prepg. ABCBA penta-block terpolymers was developed in a surfactant and org. solvent free emulsion system via one-pot RAFT polymn. The prepd. ABCBA penta-block terpolymers show large fracture tensile strain, low permanent set and potential application as elastomer materials. During the prepn. process, polyacrylic acid (PAA) was firstly synthesized as the A block via RAFT polymn. in water in the presence of a dual-functional RAFT agent. Subsequently, a partially water-sol. monomer 2,2,2-trifluoroethyl acrylate (TFEA) was used to synthesize the B block, and ABA tri-block copolymers could be obtained via org. solvent free RAFT emulsion polymn. Using the ABA tri-block copolymers as a surfactant, the ABCBA penta-block terpolymers were further synthesized using 2,2,3,4,4,4-hexafluorobutyl acrylate (HFBA) as the third monomer in an org. solvent free system. It has been found that the emulsification efficiency of the ABA tri-block copolymers is higher than that of the corresponding AB di-block copolymers. The assembly mechanism of penta-block terpolymers to form a tri-layer core-shell latex particle was proposed and confirmed by TEM images. Finally, a transparent and hydrophobic elastomeric film was obtained by directly casting the synthesized penta-block terpolymer latexes and subsequent heat treatment. The films have excellent mech. properties and water resistance.
- 34Wu, X.; Qiao, Y.; Yang, H.; Wang, J. Self-Assembly of a Series of Random Copolymers Bearing Amphiphilic Side Chains. J. Colloid Interface Sci. 2010, 349 (2), 560– 564, DOI: 10.1016/j.jcis.2010.05.09334https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXpt1Sktrw%253D&md5=f075923332cfb1412f93b0aab6d75fd5Self-assembly of a series of random copolymers bearing amphiphilic side chainsWu, Xu; Qiao, Yingjie; Yang, Hui; Wang, JinbenJournal of Colloid and Interface Science (2010), 349 (2), 560-564CODEN: JCISA5; ISSN:0021-9797. (Elsevier B.V.)A novel series of comb-like random copolymers were prepd. by polymn. of amphiphilic macromonomers, 2-(acrylamido)-octane sulfonic acid (AMC8S), 2-(acrylamido)-dodecane sulfonic acid (AMC12S), and 2-(acrylamido)-hexadecane sulfonic acid (AMC16S), with 2-(acrylamido)-2-methylpropanesulfonic acid (AMPS) resp. The synthesis of the polymers with the same contents of amphiphilic units as side chains, but different chain length, enabled us to study the chain length dependence of their assocn. in salt soln. Steady-state fluorescence measurements with pyrene as a polarity probe, quasielastic light scattering techniques (QELS) and transmission electron micrograph (TEM) were employed to investigate the associative properties of the system. The above investigations showed that all kinds of side chains begin to assemble at certain polymer concns. and the crit. aggregation concn. (CAC) decrease dramatically with the increase in the length and content of alkyl. An interesting phenomenon is that the assembly tends more favorably to occur among different mols. rather than within single mol. when the no. of carbon atoms in the alkyl groups or the polymer concn. increases, leading to the formation of larger multimol. micelle-like aggregate. The aim of the present work is to establish the fundamental preconditions of intramol. and intermol. assocn. fashions for the polymers, which is useful for the exploitation of functional groups and contributes to the development of amphiphilic random polymers.
- 35Kawata, T.; Hashidzume, A.; Sato, T. Micellar Structure of Amphiphilic Statistical Copolymers Bearing Dodecyl Hydrophobes in Aqueous Media. Macromolecules 2007, 40 (4), 1174– 1180, DOI: 10.1021/ma062299x35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXmtFSktg%253D%253D&md5=96d988e993768de20f3d0805aaedcc4cMicellar Structure of Amphiphilic Statistical Copolymers Bearing Dodecyl Hydrophobes in Aqueous MediaKawata, Takefumi; Hashidzume, Akihito; Sato, TakahiroMacromolecules (2007), 40 (4), 1174-1180CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The structure of micellar aggregates formed from ionic statistical copolymers of N-acryloyl-amino acids and N-dodecylmethacrylamide in 0.05 M aq. NaCl was studied by light scattering and fluorescence. The exptl. results indicated that the tendency of the interchain aggregation increased with the hydrophobic monomer content for each series of the copolymers but decreased with hydrophobicity of the amino acid residue in the copolymers. On the other hand, while the micellar structure of the statistical copolymers strongly depended on the kind of the hydrophobe, it is little dependent on the kind of the electrolyte monomer unit at low ionic strength. Using these data, a theor. anal. taking into account the chain stiffness effect revealed that uni-core micelles formed from the ionic statistical copolymers were of flower type with a min. loop size.
- 36Zhu, X.; Liu, M. Self-Assembly and Morphology Control of New l -Glutamic Acid-Based Amphiphilic Random Copolymers: Giant Vesicles, Vesicles, Spheres, and Honeycomb Film. Langmuir 2011, 27 (21), 12844– 12850, DOI: 10.1021/la202680j36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXht1OrtrbN&md5=84bfb08f30edb16a065a4d5e0292d0a3Self-Assembly and Morphology Control of New L-Glutamic Acid-Based Amphiphilic Random Copolymers: Giant Vesicles, Vesicles, Spheres, and Honeycomb FilmZhu, Xuewang; Liu, MinghuaLangmuir (2011), 27 (21), 12844-12850CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)New amphiphilic random copolymers contg. hydrophobic dodecyl (C12) chain and hydrophilic L-glutamic acid were synthesized, and their self-assembly in soln. as well as on the solid surfaces was investigated. The self-assembly behavior of these polymers are largely dependent on their hydrophilic and hydrophobic balances. The copolymer with a more hydrophobic alkyl chain (∼90%) self-assembled into giant vesicles with a diam. of several micrometers in a mixed solvent of ethanol and water. When the hydrophobic ratio decreased to ca. 76%, the polymer self-assembled into conventional vesicles with several hundred nanometers. The giant vesicles could be fused in certain conditions, while the conventional vesicles were stable. When the content of the hydrophilic part was further increased, no organized structures were formed. On the other hand, when the copolymer solns. were directly cast on solid substrates such as silicon plates, films with organized nanostructures could also be obtained, the morphol. of which depended on solvent selection. When ethanol or methanol was used, spheres were obtained. When dichloromethane was used as the solvent, honeycomb-like morphologies were obtained. These results showed that through appropriate mol. design, random copolymer could self-assemble into various organized structures, which could be regulated through the hydrophobic/hydrophilic balance and the solvents.
- 37Tian, F.; Yu, Y.; Wang, C.; Yang, S. Consecutive Morphological Transitions in Nanoaggregates Assembled from Amphiphilic Random Copolymer via Water Driven Micellization and Light-Triggered Dissociation. Macromolecules 2008, 41 (10), 3385– 3388, DOI: 10.1021/ma800142j37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltlalsro%253D&md5=1c16592aaf10bede6627e031b226fe94Consecutive Morphological Transitions in Nanoaggregates Assembled from Amphiphilic Random Copolymer via Water-Driven Micellization and Light-Triggered DissociationTian, Feng; Yu, Yuanyuan; Wang, Changchun; Yang, ShuMacromolecules (Washington, DC, United States) (2008), 41 (10), 3385-3388CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)We report a study of continuous morphol. transitions from spherical micelles, through hollow tubes and wormlike rods, to large vesicles in nanoaggregates self-assembled from photoresponsive amphiphilic random copolymer. Through esterification of the side chain hydroxyl groups, the hydrophilic poly(hydroxyethyl methacrylate) (PHEMA) backbone is partially and randomly modified by the hydrophobic, light-responsive 2-diazo-1,2-naphthoquinone (DNQ) mols. The micellization and transition are then triggered simply by adding water into the copolymer/DMF soln. Upon irradn. to UV light at 405 nm, large vesicles become increasingly hydrophilic with time and gradually dissoc. into globular hydrogel particles in the aq. soln. (Scheme 1), releasing the encapsulated dye mols.
- 38Sun, G.; Zhang, M.; He, J.; Ni, P. Synthesis of Amphiphilic Cationic Copolymers Poly[2-(Methacryloyloxy)Ethyl Trimethylammonium Chloride-co-stearyl Methacrylate] and Their Self-Assembly Behavior in Water and Water-Ethanol Mixtures. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 4670– 4684, DOI: 10.1002/pola.2351738https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXpvFCksb0%253D&md5=72e4ba806b4ccb2973547000d316ec1bSynthesis of amphiphilic cationic copolymers poly[2-(methacryloyloxy)ethyl trimethylammonium chloride-co-stearyl methacrylate] and their self-assembly behavior in water and water-ethanol mixturesSun, Guixiang; Zhang, Mingzu; He, Jinlin; Ni, PeihongJournal of Polymer Science, Part A: Polymer Chemistry (2009), 47 (18), 4670-4684CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)A series of amphiphilic cationic random copolymers, namely poly[2-(methacryloyloxy)ethyl trimethylammonium chloride-co-stearyl methacrylate] or poly(MADQUAT-co-SMA), were synthesized via conventional free-radical copolymn. using 2,2'-azobisisobutyronitrile (AIBN) as initiator and n-dodecanethiol as chain transfer agent. The resultant products were then characterized by FT-IR, 1H NMR, MALDI-TOF MS measurements. From the no.-av. mol. wts. of the copolymers, we can conclude that these copolymers have oligomeric structure with a limited no. of hydrophilic and hydrophobic moieties in a short polymer chain. The reactivity ratios (rMADQUAT = 0.83, rSMA = 0.25) between the hydrophilic MADQUAT monomer and the hydrophobic SMA monomer were calcd. by the Finemann and Ross method, which was based on the results of 1H NMR anal. The surface activity of the random copolymers was studied by the combination of surface tension and contact angle measurement, and these copolymers possess relatively high surface activity. The crit. aggregation concns. (cac) of the copolymers in aq. soln. were detd. by fluorescence probe method and surface tension measurement. The different nanoparticles of poly(MADQUAT-co-SMA) copolymers formed in pure water or ethanol-water mixt. were proved by the particle size and size distribution in the measurement of dynamic light scattering (DLS). Furthermore, using TEM, we could observe various self-assembly morphologies of these random copolymer. All the amphiphilic cationic random copolymers have a good self-assembly behavior, even if they are ill-defined copolymers.
- 39Stephan, T.; Muth, S.; Schmidt, M. Shape Changes of Statistical Copolymacromonomers: From Wormlike Cylinders to Horseshoe- and Meanderlike Structures. Macromolecules 2002, 35 (27), 9857– 9860, DOI: 10.1021/ma025711r39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XptFOks74%253D&md5=df799d57f47cf53452c936948493ef19Shape Changes of Statistical Copolymacromonomers: From Wormlike Cylinders to Horseshoe- and Meander-like StructuresStephan, Tim; Muth, Sandra; Schmidt, ManfredMacromolecules (2002), 35 (27), 9857-9860CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Depending on the solvent, statistical copolymer cylindrical brushes adopt different shapes from worm-like to horseshoe or meander-like structures when spin-cast onto mica. To the best of our knowledge, this is the first time that shape persistent synthetic macromols. could be chem. manipulated to change their conformation from a wormlike to regularly curved structures. It should be mentioned, though, that for cylindrical brushes with Bu acrylate side chains a collapse of the wormlike conformation to spherical structures could be induced under pressure on a Langmuir-Blodgett trough. Also, poly(p-phenylene)s with amphipolar dendritic side chains were reported to have the potential to intramol. segregate lengthwise, particularly when spread on a LB trough as well. Our next step will address the question of whether the collapse of the minority component could be detected by scattering expts. in soln., which, however, was not successful so far.
- 40Ilhan, F.; Galow, T. H.; Gray, M.; Clavier, G.; Rotello, V. M. Giant Vesicle Formation through Self-Assembly of Complementary Random Copolymers. J. Am. Chem. Soc. 2000, 122 (24), 5895– 5896, DOI: 10.1021/ja001196640https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjs1elt7Y%253D&md5=86ebd81f301c14d25ad679cf7e6098e5Giant Vesicle Formation through Self-Assembly of Complementary Random CopolymersIlhan, Faysal; Galow, Trent H.; Gray, Mark; Clavier, Gilles; Rotello, Vincent M.Journal of the American Chemical Society (2000), 122 (24), 5895-5896CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Differential interference contrast optical microscopy and confocal laser scanning microscopy showed that vesicular H-bond complexes formed between chains of 4-(chloromethyl)styrene-styrene copolymer (I)-2,6-bis(propionamido)pyrid-4-one adducts and I-thymine-1-acetic acid adducts in CHCl3 solns.
- 41Liu, X.; Kim, J.-S.; Wu, J.; Eisenberg, A. Bowl-Shaped Aggregates from the Self-Assembly of an Amphiphilic Random Copolymer of Poly(Styrene- Co -Methacrylic Acid). Macromolecules 2005, 38 (16), 6749– 6751, DOI: 10.1021/ma050665r41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmt1Krtrw%253D&md5=3d6e4ea1e6486b954e76bbc45a4b6fa2Bowl-Shaped Aggregates from the Self-Assembly of an Amphiphilic Random Copolymer of Poly(styrene-co-methacrylic acid)Liu, Xiaoya; Kim, Joon-Seop; Wu, Jun; Eisenberg, AdiMacromolecules (2005), 38 (16), 6749-6751CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A bowl-shaped structure of random poly(styrene-co-methacrylic acid) was prepd. by self-assembly in dil. soln. We emphasize that the copolymer was a random copolymer in contrast to the block copolymers or functionally terminated homopolymers used before. The added viscosity control is probably provided by H bond interactions among OH groups along the backbone as in encountered in carboxylic acids in bulk or in org. solvents.
- 42Neal, T. J.; Parnell, A. J.; King, S. M.; Beattie, D. L.; Murray, M. W.; Williams, N. S. J.; Emmett, S. N.; Armes, S. P.; Spain, S. G.; Mykhaylyk, O. O. Control of Particle Size in the Self-Assembly of Amphiphilic Statistical Copolymers. Macromolecules 2021, 54 (3), 1425– 1440, DOI: 10.1021/acs.macromol.0c0234142https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhslehs7o%253D&md5=a59483ff3e5c24287beb7c07816f64dcControl of Particle Size in the Self-Assembly of Amphiphilic Statistical CopolymersNeal, Thomas J.; Parnell, Andrew J.; King, Stephen M.; Beattie, Deborah L.; Murray, Martin W.; Williams, Neal S. J.; Emmett, Simon N.; Armes, Steven P.; Spain, Sebastian G.; Mykhaylyk, Oleksandr O.Macromolecules (Washington, DC, United States) (2021), 54 (3), 1425-1440CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A range of amphiphilic statistical copolymers is synthesized where the hydrophilic component is either methacrylic acid (MAA) or 2-(dimethylamino)ethyl methacrylate (DMAEMA) and the hydrophobic component comprises Me, Et, Bu, hexyl, or 2-ethylhexyl methacrylate, which provide a broad range of partition coeffs. (log P). Small-angle x-ray scattering studies confirm that these amphiphilic copolymers self-assemble to form well-defined spherical nanoparticles in an aq. soln., with more hydrophobic copolymers forming larger nanoparticles. Varying the nature of the alkyl substituent also influenced self-assembly with more hydrophobic comonomers producing larger nanoparticles at a given copolymer compn. A model based on particle surface charge d. (PSC model) is used to describe the relation between copolymer compn. and nanoparticle size. This model assumes that the hydrophilic monomer is preferentially located at the particle surface and provides a good fit to all of the exptl. data. More specifically, a linear relation is obsd. between the surface area fraction covered by the hydrophilic comonomer required to achieve stabilization and the log P value for the hydrophobic comonomer. Contrast variation small-angle neutron scattering is used to study the internal structure of these nanoparticles. This technique indicates partial phase sepn. within the nanoparticles, with about half of the available hydrophilic comonomer repeat units being located at the surface and hydrophobic comonomer-rich cores. This information enables a refined PSC model to be developed, which indicates the same relation between the surface area fraction of the hydrophilic comonomer and the log P of the hydrophobic comonomer repeat units for the anionic (MAA) and cationic (DMAEMA) comonomer systems. This study demonstrates how nanoparticle size can be readily controlled and predicted using relatively ill-defined statistical copolymers, making such systems a viable attractive alternative to diblock copolymer nanoparticles for a range of industrial applications.
- 43Azuma, Y.; Terashima, T.; Sawamoto, M. Self-Folding Polymer Iron Catalysts for Living Radical Polymerization. ACS Macro Lett. 2017, 6 (8), 830– 835, DOI: 10.1021/acsmacrolett.7b0049843https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhtF2hurnN&md5=57e0d5b1b3ad8706735d4be8edc6faf7Self-Folding Polymer Iron Catalysts for Living Radical PolymerizationAzuma, Yusuke; Terashima, Takaya; Sawamoto, MitsuoACS Macro Letters (2017), 6 (8), 830-835CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)Iron-bearing self-folding polymers were created with amphiphilic random copolymers as active, versatile, and recyclable polymer-supported catalysts for living radical polymn. (LRP). The key is to build bis(imino)pyridine ligand cavities for iron complexes as linking units within self-folding polymers. Self-folding polymer ligands are synthesized by the intramol. imine crosslinking of self-folded amphiphilic random copolymers bearing hydrophilic poly(ethylene glycol) (PEG), hydrophobic dodecyl, and urea/aniline pendants with 2,6-pyridinedicarboxaldehyde in water. The folding polymers efficiently formed iron complex catalysts in the cores to induce LRP and random or block copolymn. of various methacrylates. The self-folding polymer catalysts not only showed high activity and tolerance to functional groups such as acid, hydroxyl groups, and oxygen but also afforded easy product recovery and catalyst recycle thanks to hydrophilic PEG chains.
- 44Lyon, C. K.; Prasher, A.; Hanlon, A. M.; Tuten, B. T.; Tooley, C. A.; Frank, P. G.; Berda, E. B. A Brief User’s Guide to Single-Chain Nanoparticles. Polym. Chem. 2015, 6 (2), 181– 197, DOI: 10.1039/C4PY01217H44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVKmtbrP&md5=6018dc5ae1b7d54385638a6da6e6f4c2A brief user's guide to single-chain nanoparticlesLyon, Christopher K.; Prasher, Alka; Hanlon, Ashley M.; Tuten, Bryan T.; Tooley, Christian A.; Frank, Peter G.; Berda, Erik B.Polymer Chemistry (2015), 6 (2), 181-197CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)In this review we outline the various methods that have been explored to synthesize architecturally defined nanoparticles from discrete polymer chains, summarize the methods of characterization that are required to prove their formation and probe their morphol., and introduce a no. of potential applications that are being explored currently. Given the small size of the nanostructures produced by these methods and the relative ease with which they can be tailored to specific end use applications it is likely such efforts will intensify in the coming years. So far, simple chem. has been utilized and high-level characterization and modeling studies have been applied to understand the process by which these particles form and how they behave, both in the bulk and in soln. Although impossible to predict where this work will lead, we hope this "user's guide" will prove useful to the community as research on single-chain nanoparticles continues to evolve.
- 45Romulus, J.; Weck, M. Single-Chain Polymer Self-Assembly Using Complementary Hydrogen Bonding Units. Macromol. Rapid Commun. 2013, 34 (19), 1518– 1523, DOI: 10.1002/marc.20130050145https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlSiu7bP&md5=f2a6a6fa1b58d398fc781d53ecebab8dSingle-Chain Polymer Self-Assembly Using Complementary Hydrogen Bonding UnitsRomulus, Joy; Weck, MarcusMacromolecular Rapid Communications (2013), 34 (19), 1518-1523CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)A triblock copolymer contg. the complementary hydrogen bonding recognition pair ureidoguanosine-diaminonaphthyridine (UG-DAN) as pendant functional groups is synthesized using ring-opening metathesis polymn. (ROMP). The norbornene-based DAN monomer is shown to allow for a controlled polymn. when polymd. in the presence of a modified-UG mol. that serves as a protecting group, subsequently allowing for the fabrication of functionalized triblock copolymers. The self-assembly of the copolymers was characterized using dynamic light scattering and 1H NMR spectroscopy. It is demonstrated that the polymers self-assemble via complementary hydrogen bonding motifs even at low dilns., indicating intramol. interactions.
- 46Gillissen, M. A. J.; Voets, I. K.; Meijer, E. W.; Palmans, A. R. A. Single Chain Polymeric Nanoparticles as Compartmentalised Sensors for Metal Ions. Polym. Chem. 2012, 3 (11), 3166– 3174, DOI: 10.1039/c2py20350b46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVGisbvJ&md5=546ada39794d7cbb342ec917911bab5fSingle chain polymeric nanoparticles as compartmentalised sensors for metal ionsGillissen, Martijn A. J.; Voets, Ilja K.; Meijer, E. W.; Palmans, Anja. R. A.Polymer Chemistry (2012), 3 (11), 3166-3174CODEN: PCOHC2; ISSN:1759-9962. (Royal Society of Chemistry)3,3'-Bis(acylamino)-2,2'-bipyridine substituted benzene-1,3,5-tricarboxamide (BiPy-BTA) grafted polynorbornene polymers were prepd. via ring-opening metathesis polymn. using a third generation Grubbs catalyst. The polymers fold intramolecularly via π-π interactions into fluorescent, compartmentalized particles of nanometer-size in mixts. of THF and methylcyclohexane. Spectroscopic and light scattering techniques show that the compact conformation of the folded polymer is affected by increasing the BiPy-BTA functionalization degree and by changing the solvent polarity. Changes in the conformation are accompanied by changes in the fluorescence intensity. Due to the affinity of the 2,2'-bipyridine units for metal ions such as Cu, the particles obtained are effective sensors for these metals. The compartmentalization of the binding motifs in SCPNs proves to be advantageous in sensor applications of these particles.
- 47Terashima, T.; Sugita, T.; Fukae, K.; Sawamoto, M. Synthesis and Single-Chain Folding of Amphiphilic Random Copolymers in Water. Macromolecules 2014, 47 (2), 589– 600, DOI: 10.1021/ma402355v47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXnsFersg%253D%253D&md5=05f57b8e32e1a5aa4f9f830e7ff5e3ddSynthesis and Single-Chain Folding of Amphiphilic Random Copolymers in WaterTerashima, Takaya; Sugita, Takanori; Fukae, Kaoru; Sawamoto, MitsuoMacromolecules (Washington, DC, United States) (2014), 47 (2), 589-600CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Amphiphilic random methacrylate copolymers, consisting of poly(ethylene glycol) (PEG) and alkyl pendent groups, undergo reversible single-chain self-folding in water via intramol. hydrophobic interaction, to generate a dynamic unimol. hydrophobic nanospace, similar in shape but structurally different relative to micelles and microgel star polymers. These copolymers were prepd. by the ruthenium-catalyzed living radical copolymn. of a PEG methacrylate (PEGMA) and an alkyl methacrylate (RMA; R, -CnH2n+1, n = 1-18), where copolymer compn., d.p., and hydrophobic R moiety were varied. Detailed structural and chain-folding characterization has revealed: single-chain folding is favored with the RMA content 20-40 mol % per chain; the hydrophobic inner compartment (or the self-folded structure) is stable even at a high polymer concn. (up to ∼6 wt. %); and folded-unfolded transition occurs on addn. of methanol or by elevating soln. temp., finally to phase-sepn. above a lower crit. soln. temp.
- 48Terashima, T.; Mes, T.; De Greef, T. F. A.; Gillissen, M. A. J.; Besenius, P.; Palmans, A. R. A.; Meijer, E. W. Single-Chain Folding of Polymers for Catalytic Systems in Water. J. Am. Chem. Soc. 2011, 133 (13), 4742– 4745, DOI: 10.1021/ja200449448https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjtFeku70%253D&md5=a4f7c01b74384b16c8f00c7b2ab04641Single-Chain Folding of Polymers for Catalytic Systems in WaterTerashima, Takaya; Mes, Tristan; De Greef, Tom F. A.; Gillissen, Martijn A. J.; Besenius, Pol; Palmans, Anja R. A.; Meijer, E. W.Journal of the American Chemical Society (2011), 133 (13), 4742-4745CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Enzymes are a source of inspiration for chemists attempting to create versatile synthetic catalysts. In order to arrive at a polymeric chain carrying catalytic units sepd. spatially, it is a prerequisite to fold these polymers in water into well-defined compartmentalized architectures thus creating a catalytic core. Herein, we report the synthesis, phys. properties, and catalytic activity of a water-sol. segmented terpolymer in which a helical structure in the apolar core is created around a ruthenium-based catalyst. The supramol. chirality of this catalytic system is the result of the self-assembly of benzene-1,3,5-tricarboxamide side chains, while the catalyst arises from the sequential ruthenium-catalyzed living radical polymn. of the different monomers followed by ligand exchange. The polymers exhibit a two-state folding process and show transfer hydrogenation in water.
- 49Matsumoto, M.; Sawamoto, M.; Terashima, T. Orthogonal Folding of Amphiphilic/Fluorous Random Block Copolymers for Double and Multicompartment Micelles in Water. ACS Macro Lett. 2019, 8 (3), 320– 325, DOI: 10.1021/acsmacrolett.9b0007849https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXkt1eisb4%253D&md5=bf44aff6acc6c9b7b9f35cbf970b1d9fOrthogonal Folding of Amphiphilic/Fluorous Random Block Copolymers for Double and Multicompartment Micelles in WaterMatsumoto, Mayuko; Sawamoto, Mitsuo; Terashima, TakayaACS Macro Letters (2019), 8 (3), 320-325CODEN: AMLCCD; ISSN:2161-1653. (American Chemical Society)Here, we report orthogonal folding and self-assembly systems of amphiphilic/fluorous random block copolymers for double core and multicompartment micelles in water. For this, we developed the precision folding techniques of polymer chains via the selective self-assembly of the pendant groups. Typically, A/C-B/C random block copolymers were designed: Hydrophobic dodecyl groups (A) and fluorous fluorinated octyl groups (B) were introduced into the resp. blocks, while hydrophilic poly(ethylene glycol) chains (C) were randomly incorporated into all the segments. By controlling the chain length and compn. of the resp. blocks, the copolymers induce orthogonal single-chain folding in water to form double-compartment micelles comprising hydrophobic and fluorous cores. The copolymers were site-selectively folded in a fluoroalc. to result in tadpole unimer micelles comprising a hydrophobic A/C unimer micelle and an unfolded fluorous B/C chain. Addnl., asym. A/C-B/C random block copolymers with short and highly hydrophobic or fluorous segments were effective for multicompartment micelles in water.
- 50Matsumoto, M.; Terashima, T.; Matsumoto, K.; Takenaka, M.; Sawamoto, M. Compartmentalization Technologies via Self-Assembly and Cross-Linking of Amphiphilic Random Block Copolymers in Water. J. Am. Chem. Soc. 2017, 139 (21), 7164– 7167, DOI: 10.1021/jacs.7b0315250https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXnslKis7k%253D&md5=5de60e0783e388ccc071895904f7af55Compartmentalization Technologies via Self-Assembly and Cross-Linking of Amphiphilic Random Block Copolymers in WaterMatsumoto, Mayuko; Terashima, Takaya; Matsumoto, Kazuma; Takenaka, Mikihito; Sawamoto, MitsuoJournal of the American Chemical Society (2017), 139 (21), 7164-7167CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Orthogonal self-assembly and intramol. crosslinking of amphiphilic random block copolymers in water afforded an approach to tailor-make well-defined compartments and domains in single polymer chains and nanoaggregates. For a double compartment single-chain polymer, an amphiphilic random block copolymer bearing hydrophilic poly(ethylene glycol) (PEG) and hydrophobic dodecyl, benzyl, and olefin pendants was synthesized by living radical polymn. (LRP) and postfunctionalization; the dodecyl and benzyl units were incorporated into the different block segments, whereas PEG pendants were statistically attached along a chain. The copolymer self-folded via the orthogonal self-assembly of hydrophobic dodecyl and benzyl pendants in water, followed by intramol. crosslinking, to form a single-chain polymer carrying double yet distinct hydrophobic nanocompartments. A single-chain crosslinked polymer with a chlorine terminal served as a globular macroinitiator for LRP to provide an amphiphilic tadpole macromol. comprising a hydrophilic nanoparticle and a hydrophobic polymer tail; the tadpole thus self-assembled into multicompartment aggregates in water.
- 51Neal, T. J.; Beattie, D. L.; Byard, S. J.; Smith, G. N.; Murray, M. W.; Williams, N. S. J.; Emmett, S. N.; Armes, S. P.; Spain, S. G.; Mykhaylyk, O. O. Self-Assembly of Amphiphilic Statistical Copolymers and Their Aqueous Rheological Properties. Macromolecules 2018, 51 (4), 1474– 1487, DOI: 10.1021/acs.macromol.7b0213451https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitlarsrw%253D&md5=a06b2c97ac88e4a06405b09b3e1dcdd3Self-Assembly of Amphiphilic Statistical Copolymers and Their Aqueous Rheological PropertiesNeal, Thomas J.; Beattie, Deborah L.; Byard, Sarah J.; Smith, Gregory N.; Murray, Martin W.; Williams, Neal S. J.; Emmett, Simon N.; Armes, Steven P.; Spain, Sebastian G.; Mykhaylyk, Oleksandr O.Macromolecules (Washington, DC, United States) (2018), 51 (4), 1474-1487CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A range of poly(Bu methacrylate-stat-methacrylic acid) [P(BMA-stat-MAA)] statistical copolymers of various compns. and mol. wts. ranging from 5 to 30 kDa were prepd. using either reversible addn.-fragmentation chain transfer (RAFT) soln. copolymn. or conventional free radical polymn. in isopropanol (IPA). On diln. with water, these amphiphilic copolymers self-assembled to form spherical nano-objects as confirmed by small-angle X-ray scattering (SAXS) and transmission electron microscopy. Various structural models were examd. to ext. information regarding the mean nano-object size and morphol. It is found that nano-object radii are independent of copolymer mol. wt. but depend on the copolymer compn.: the smaller the amt. of MAA units in the copolymer chains, the larger the nano-objects that are formed. Combined SAXS and aq. electrophoretic measurements indicated that most of the MAA units are located at the nano-object surface. Furthermore, SAXS and rheol. measurements were used to monitor the effect of solvent compn. on the copolymer morphol. both at a fixed copolymer concn. (either 1 or 25 wt. %) and also for a gradual variation in copolymer concns. (from 1 to 40 wt. %) when adding water to the initial copolymer soln. in IPA. These studies revealed that the copolymers are present in soln. as molecularly dissolved Gaussian chains when the solvent compn. is IPA-rich. However, the copolymer chains self-assemble into spherical nano-objects when the solvent compn. is water-rich. At intermediate solvent compns., SAXS anal. confirmed the formation of an interconnected nano-object network, which accounts for the apparently anomalous increase in soln. viscosity on diln. indicated by rheol. measurements.
- 52Hattori, G.; Takenaka, M.; Sawamoto, M.; Terashima, T. Nanostructured Materials via the Pendant Self-Assembly of Amphiphilic Crystalline Random Copolymers. J. Am. Chem. Soc. 2018, 140 (27), 8376– 8379, DOI: 10.1021/jacs.8b0383852https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtFCms77J&md5=06ffde6a8667e2617a6a6ab1b0aabcdeNanostructured Materials via the Pendant Self-Assembly of Amphiphilic Crystalline Random CopolymersHattori, Goki; Takenaka, Mikihito; Sawamoto, Mitsuo; Terashima, TakayaJournal of the American Chemical Society (2018), 140 (27), 8376-8379CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Versatile self-assembly systems to nanostructured materials in both solid and soln. were developed with common amphiphilic random copolymers bearing hydrophilic poly(ethylene glycol) (PEG) and hydrophobic cryst. octadecyl pendants. The copolymers efficiently induced precision self-assembly of the pendants to provide not only core-cryst., thermoresponsive micelles and vesicles in water and reverse micelles in hexane but also sub-10 nm lamellar or spherical microphase sepn. structure in solid. Typically, the solid random copolymers with 50-80 mol % octadecyl units formed lamellar structure of a hydrophilic PEG layer and a hydrophobic, cryst. octadecyl layer. Importantly, the domain spacing is about 5 nm, much smaller than that generally obtained with conventional block copolymers. The domain structure is controlled by compn., independent of chain length. The copolymers further gave various thermoresponsive, compartmentalized materials in aq. and org. media, where the 3D structure can be also controlled by the compn. and sample prepn. protocols.
- 53Ilavsky, J.; Jemian, P. R. Irena : Tool Suite for Modeling and Analysis of Small-Angle Scattering. J. Appl. Crystallogr. 2009, 42 (2), 347– 353, DOI: 10.1107/S002188980900222253https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXjsFSnsbY%253D&md5=ebf014363d395cb4180fe0fd96f7f714Irena: tool suite for modeling and analysis of small-angle scatteringIlavsky, Jan; Jemian, Peter R.Journal of Applied Crystallography (2009), 42 (2), 347-353CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Irena, a tool suite for anal. of both x-ray and neutron small-angle scattering (SAS) data within the com. Igor Pro application, brings together a comprehensive suite of tools useful for studies in materials science, physics, chem., polymer science and other fields. In addn. to Guinier and Porod fits, the suite combines a variety of advanced SAS data evaluation tools for the modeling of size distribution in the dil. limit using max. entropy and other methods, dil. limit small-angle scattering from multiple noninteracting populations of scatterers, the pair-distance distribution function, a unified fit, the Debye-Bueche model, the reflectivity (x-ray and neutron) using Parratt's formalism, and small-angle diffraction. There are also a no. of support tools, such as a data import/export tool supporting a broad sampling of common data formats, a data modification tool, a presentation-quality graphics tool optimized for small-angle scattering data, and a neutron and x-ray scattering contrast calculator. These tools are brought together into one suite with consistent interfaces and functionality. The suite allows robust automated note recording and saving of parameters during export.
- 54Horcas, I.; Fernández, R.; Gómez-Rodríguez, J. M.; Colchero, J.; Gómez-Herrero, J.; Baro, A. M. WSXM: A Software for Scanning Probe Microscopy and a Tool for Nanotechnology. Rev. Sci. Instrum. 2007, 78, 013705, DOI: 10.1063/1.243241054https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXis1Srsrg%253D&md5=ff73289035c4652ebb9800e362892410WSXM: a software for scanning probe microscopy and a tool for nanotechnologyHorcas, I.; Fernandez, R.; Gomez-Rodriguez, J. M.; Colchero, J.; Gomez-Herrero, J.; Baro, A. M.Review of Scientific Instruments (2007), 78 (1), 013705/1-013705/8CODEN: RSINAK; ISSN:0034-6748. (American Institute of Physics)In this work we briefly describe the most relevant features of WSXM, a freeware scanning probe microscopy software based on MS-Windows. The article is structured in three different sections: The introduction is a perspective on the importance of software on scanning probe microscopy. The second section is devoted to describe the general structure of the application; in this section the capabilities of WSXM to read third party files are stressed. Finally, a detailed discussion of some relevant procedures of the software is carried out.
- 55Kheshgi, S.; Scriven, L. E. Dewetting: Nucleation and Growth of Dry Regions. Chem. Eng. Sci. 1991, 46 (2), 519– 526, DOI: 10.1016/0009-2509(91)80012-N55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhtVGjs7k%253D&md5=bd33a4450b0cef9f1c03355fbd9dbf60Dewetting: nucleation and growth of dry regionsKheshgi, Haroon S.; Scriven, L. E.Chemical Engineering Science (1991), 46 (2), 519-26CODEN: CESCAC; ISSN:0009-2509.Dewetting of a solid surface covered by a film of nonwetting liq. proceeds from a preexisting dry patch or edge or initiates from some film-thinning disturbance that grows until the film ruptures. Local thinning can be caused by evapn.; by drainage due to gravity or capillarity-driven flow, esp. from sharp surfaces; or by surface tension gradients, such as are caused by surfactants delivered by particles falling on the film. Once a nonwetting film is sufficiently thinned, conjoining (neg. disjoining) pressure can accelerate thinning until rupture. This catastrophic rupture is modeled by solving the Navier-Stokes system approximated for thickness variations over distances that are long compared with the mean film thickness, and augmented with conjoining pressure. Rupture leads to film retraction and formation of a dry patch. These phenomena are visualized via moire topog. Of special interest are local spreading disturbances, where airborne particles fall on the film surface; craters or dry patches often nucleate. Implications for coating operations are discussed.
- 56Xue, L.; Han, Y. Inhibition of Dewetting of Thin Polymer Films. Prog. Mater. Sci. 2012, 57 (6), 947– 979, DOI: 10.1016/j.pmatsci.2012.01.00356https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xls1Siu70%253D&md5=0c23ab0ed2e6414db3bffb2e3b764a71Inhibition of dewetting of thin polymer filmsXue, Longjian; Han, YanchunProgress in Materials Science (2012), 57 (6), 947-979CODEN: PRMSAQ; ISSN:0079-6425. (Elsevier Ltd.)A review. Strategies for the inhibition of dewetting of thin polymer films are reviewed. First, a brief introduction to the theory and the dynamics of dewetting of thin polymer films is given. Methods for the inhibition of dewetting of thin polymer films, including the modification of the substrate and the polymer, the crosslinking of the polymer and the addn. of an additive as well as their mechanisms for inhibition are discussed. The chem. modification of the substrate or the polymer itself and the phase sepn. increase the thermodn. stability of the system, while the crosslinking of the polymers reduces the mobility of the polymer chains, kinetically arresting the dewetting. The addn. of an additive appears to thermodynamically and kinetically inhibit the dewetting. Finally, the future outlook in this research field is highlighted.
- 57Barsbay, M.; Güven, O.; Stenzel, M. H.; Davis, T. P.; Barner-Kowollik, C.; Barner, L. Verification of Controlled Grafting of Styrene from Cellulose via Radiation-Induced RAFT Polymerization. Macromolecules 2007, 40 (20), 7140– 7147, DOI: 10.1021/ma070825u57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtVSgs7bF&md5=7686cbcffa76684773b4600319dc3594Verification of Controlled Grafting of Styrene from Cellulose via Radiation-Induced RAFT PolymerizationBarsbay, Murat; Gueven, Olgun; Stenzel, Martina H.; Davis, Thomas P.; Barner-Kowollik, Christopher; Barner, LeonieMacromolecules (Washington, DC, United States) (2007), 40 (20), 7140-7147CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Reversible addn.-fragmentation chain transfer (RAFT) polymn. was applied to radiation-induced graft polymn. of styrene from cellulose. The grafting of styrene from cellulose substrates using the chain transfer agent cumyl phenyldithioacetate was confirmed by Raman and XPS, differential scanning calorimetery, thermogravimetric anal., SEM, and contact angle anal. Grafted polystyrene chains were cleaved from the cellulose surface by acidic hydrolysis of the cellulose. The no.-av. mol. wt. and polydispersity index of the grafted and the free (nongrafted) polystyrenes obtained under identical conditions were detd. by size exclusion chromatog. Grafted and nongrafted polystyrenes have almost the same (near theor.) mol. wt. and narrow polydispersity, thus proving for the first time the control of the grafting process mediated via RAFT without any prior functionalization of the surface.
- 58Eisenberg, A.; Yokoyama, T.; Sambalido, E. Dehydration Kinetics and Glass Transition of Poly(Acrylic Acid). J. Polym. Sci. Part A-1 Polym. Chem. 1969, 7 (7), 1717– 1728, DOI: 10.1002/pol.1969.15007071458https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXksVartQ%253D%253D&md5=7168b32a3f0f9c9b61cb323407e90d32Dehydration kinetics and glass transition of poly(acrylic acid)Eisenberg, Adi; Yokoyama, Tetsuo; Sambalido, EmmaJournal of Polymer Science, Part A-1: Polymer Chemistry (1969), 7 (7), 1717-28CODEN: JPSPC3; ISSN:0449-296X.The kinetics of dehydration and decarboxylation as well as the glass transition temp. as a function of anhydride content were measured for poly(acrylic acid) (I). The glass transition of I is of the order of 103° and increases with increasing anhydride content, reaching an extrapolated value of 140° for the pure linear anhydride. Anhydride formation is a first-order reaction, as is also decarboxylation, the latter being much slower than the former. The rate consts. are: for dehydration, ka = 2.5 × 109 exp (-26,000/RT); for decarboxylation, kd = 2.9 × 108 exp (-27,000/RT). Anhydride formation occurs primarily by an intramol. process.
- 59Chapin, E. C.; Ham, G. E.; Mills, C. L. Copolymerization. VII. Relative Rates of Addition of Various Monomers in Copolymerization. J. Polym. Sci. 1949, 4, 597– 604, DOI: 10.1002/pol.1949.12004050559https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG3cXptVOn&md5=e78bf52fb289122c0952eccccb62a5efCopolymerization. VII. Relative rates of addition of various monomers in copolymerizationChapin, Earl C.; Ham, George E.; Mills, Charles L.Journal of Polymer Science (1949), 4 (), 597-604CODEN: JPSCAU; ISSN:0022-3832.cf. C.A. 43, 8736g; 42, 4151i. Data are presented on monomer reactivity in copolymerization for these 13 systems: styrene (I)-methacrylic acid (II), I-acrylic acid (III), vinylidene chloride (IV)-crotonic acid (V), vinyl acetate (VI)-V, acrylonitrile (VII)-vinyl 2-ethylhexoate (VIII), VII-vinyl formate (IX), VII-vinyl benzoate (X), VII-α-acetoxystyrene (XI), VII-allyl chloride (XII), IV-X, IV-methyl isopropenyl ketone (XIII), vinyl chloride (XIV)-vinyl isobutyl ether (XV), and I-citraconic anydride (XVI). The reactivity ratios, γ1 and γ2, for these copolymerizations were obtained by choosing values which gave the best curve fit on substitution in the differential copolymer equation. The azeotrope of the system I-II occurs at 74% II while that of I-III occurs at 48% III. By using the Price and Alfrey relationship for the consts. Q and e these values are interpreted as evidence of higher resonance stability of the radical adduct formed from II than from III and lower polarity of II than III. Azeotropes also occur in the systems VI-V, VII-XI, and I-XVI. The copolymerization of I and XVI yielded a copolymerization curve similar to that of I-maleic anhydride reported by Alfrey and Lavin (C.A. 40, 1149.2); evidently XVI is less reactive toward a I-free radical than maleic anhydride.
- 60Mayadunne, R. T. A.; Rizzardo, E.; Chiefari, J.; Krstina, J.; Moad, G.; Postma, A.; Thang, S. H. Radical Polymerization in Two Steps. Macromolecules 2000, 33, 243– 245, DOI: 10.1021/ma991451a60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhtlWgtA%253D%253D&md5=a0ad88898f76e3c4b5eca06698d39614Living polymers by the use of trithiocarbonates as reversible addition-fragmentation chain transfer (RAFT) agents. ABA triblock copolymers by radical polymerization in two stepsMayadunne, Roshan T. A.; Rizzardo, Ezio; Chiefari, John; Krstina, Julia; Moad, Graeme; Postma, Almar; Thang, San H.Macromolecules (2000), 33 (2), 243-245CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Trithiocarbonates RS-CS-SR', where R and R' are homolytic leaving groups, were used as effective reversible addn.-fragmentation chain transfer agents. Styrene, Me acrylate, Bu acrylate and Me methacrylate were polymd. in the presence of various trithiocarbonate compds. (R/R': Me/C(Me)2CN, Me/CH(Ph)COOH, benzyl/benzyl, and α-methylbenzyl/α-methylbenzyl) and mol. wt. and conversion data were detd. Homopolymers and block copolymers of controlled mol. wt. and narrow polydispersity were prepd. Important ABA triblock copolymers of predetd. mol. wt. are accessible in 2 steps.
- 61Buzin, A. I.; Pyda, M.; Costanzo, P.; Matyjaszewski, K.; Wunderlich, B. Calorimetric Study of Block-Copolymers of Poly(n-Butyl Acrylate) and Gradient Poly(n-Butyl Acrylate-Co-Methyl Methacrylate). Polymer (Guildf). 2002, 43 (20), 5563– 5569, DOI: 10.1016/S0032-3861(02)00358-0There is no corresponding record for this reference.
- 62Sanchez-Fernandez, A.; Edler, K. J.; Arnold, T.; Heenan, R. K.; Porcar, L.; Terrill, N. J.; Terry, A. E.; Jackson, A. J. Micelle Structure in a Deep Eutectic Solvent: A Small-Angle Scattering Study. Phys. Chem. Chem. Phys. 2016, 18 (20), 14063– 14073, DOI: 10.1039/C6CP01757F62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XmvFOlsLc%253D&md5=a46225f5c8786103c0dcef785859f328Micelle structure in a deep eutectic solvent: a small-angle scattering studySanchez-Fernandez, A.; Edler, K. J.; Arnold, T.; Heenan, R. K.; Porcar, L.; Terrill, N. J.; Terry, A. E.; Jackson, A. J.Physical Chemistry Chemical Physics (2016), 18 (20), 14063-14073CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)In recent years many studies into green solvents have been undertaken and deep eutectic solvents (DES) have emerged as sustainable and green alternatives to conventional solvents since they may be formed from cheap non-toxic org. precursors. In this study we examine amphiphile behavior in these novel media to test our understanding of amphiphile self-assembly within environments that have an intermediate polarity between polar and non-polar extremes. We have built on our recently published results to present a more detailed structural characterization of micelles of sodium dodecylsulfate (SDS) within the eutectic mixt. of choline chloride and urea. Here we show that SDS adopts an unusual cylindrical aggregate morphol., unlike that seen in water and other polar solvents. A new morphol. transition to shorter aggregates was found with increasing concn. The self-assembly of SDS was also investigated in the presence of water; which promotes the formation of shorter aggregates.
- 63Gaillard, N.; Guyot, A.; Claverie, J. Block Copolymers of Acrylic Acid and Butyl Acrylate Prepared by Reversible Addition-Fragmentation Chain Transfer Polymerization: Synthesis, Characterization, and Use in Emulsion Polymerization. J. Polym. Sci. Part A Polym. Chem. 2003, 41 (5), 684– 698, DOI: 10.1002/pola.1060663https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhtlCnur0%253D&md5=d40c44d67105fc9ec9ea57155237b4feBlock copolymers of acrylic acid and butyl acrylate prepared by reversible addition-fragmentation chain transfer polymerization: synthesis, characterization, and use in emulsion polymerizationGaillard, Nicolas; Guyot, Alain; Claverie, JeromeJournal of Polymer Science, Part A: Polymer Chemistry (2003), 41 (5), 684-698CODEN: JPACEC; ISSN:0887-624X. (John Wiley & Sons, Inc.)Amphiphilic block copolymers of poly(acrylic acid-b-Bu acrylate) (I) were prepd. by reversible addn.-fragmentation chain transfer polymn. in a one-pot reaction. I were characterized by NMR, static and dynamic light scattering, tensiometry, and size exclusion chromatog. The aggregation characteristics of I corresponded to those theor. predicted for a star micelle. In a Bu acrylate-Me methacrylate emulsion copolymn., low amts. of I could stabilize the latexes with solid contents up to 50%.
- 64Chernikova, E. V.; Lysenko, E. A.; Serkhacheva, N. S.; Prokopov, N. I. Self-Assembly of Amphiphilic Block Copolymers during Reversible Addition-Fragmentation Chain Transfer Heterophase Polymerization: Problems, Achievements, and Outlook. Polym. Sci. - Ser. C 2018, 60 (2), 192– 218, DOI: 10.1134/S181123821802004264https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitlKmtbnI&md5=06f47b161ab97bcc2458ee70e63796dfSelf-Assembly of Amphiphilic Block Copolymers during Reversible Addition-Fragmentation Chain Transfer Heterophase Polymerization: Problems, Achievements, and OutlookChernikova, E. V.; Lysenko, E. A.; Serkhacheva, N. S.; Prokopov, N. I.Polymer Science, Series C: Selected Topics (2018), 60 (S1), 192-218CODEN: PSSCCH; ISSN:1555-614X. (Pleiades Publishing, Ltd.)The formation of dispersions of amphiphilic block copolymer particles with the controlled morphol. via heterophase polymn. mediated by reversible addn.-fragmentation agents is considered. Variants of dispersion, emulsion, and seeded polymns. are analyzed, and the mechanism of this process is discussed. Special attention is focused on issues related to control over the morphol. of the formed particles immediately during synthesis and methods of its transformation in the resulting dispersions.
- 65Chenal, M.; Rieger, J.; Vechambre, C.; Chenal, J.-M.; Chazeau, L.; Creton, C.; Bouteiller, L. Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard Phase. Macromol. Rapid Commun. 2013, 34, 1524– 1529, DOI: 10.1002/marc.20130041065https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtlahtLbO&md5=4d4bc64299888278198f8cdf39b17d57Soft Nanostructured Films with an Ultra-Low Volume Fraction of Percolating Hard PhaseChenal, Marion; Rieger, Jutta; Vechambre, Cyril; Chenal, Jean-Marc; Chazeau, Laurent; Creton, Costantino; Bouteiller, LaurentMacromolecular Rapid Communications (2013), 34 (19), 1524-1529CODEN: MRCOE3; ISSN:1022-1336. (Wiley-VCH Verlag GmbH & Co. KGaA)In this study, aq. emulsion polymn. of Bu acrylate is performed in batch conditions without surfactants using a poly(acrylic acid)-trithiocarbonate macro-RAFT agent to control the polymn. and to stabilize the emulsion. According to the polymn.-induced self-assembly (PISA) approach, well-defined amphiphilic PAA-b-PBA diblock copolymers form and self-assemble during synthesis to yield highly stable core-shell particles with an extremely thin hard PAA shell. For the first time, we report here the specific properties of films obtained from these particular latexes. After drying the aq. dispersion, tough and transparent films are obtained. Although the films are not chem. crosslinked, they do not dissolve in good solvents for PBA. Moreover, they remain transparent even after immersion in water. Rheol. shows that the films are both stiff and ductile, thanks to the nanostructured but very low vol. fraction (less than 3 wt%) of PAA forming a percolating network in the soft PBA. Compared with conventional core-shell-based films, this approach affords for the first time a route to a thin percolating honeycomb nanostructure with a sharp and strong interface between the two phases. The versatility of the synthetic procedure opens perspectives for a large range of functional materials.
- 66Pedersen, J. S.; Gerstenberg, M. C. Scattering Form Factor of Block Copolymer Micelles. Macromolecules 1996, 29 (4), 1363– 1365, DOI: 10.1021/ma951211566https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XltlWjsA%253D%253D&md5=e6f549fc773b455a50c0d6e157ca4981Scattering Form Factor of Block Copolymer MicellesPedersen, Jan Skov; Gerstenberg, Michael C.Macromolecules (1996), 29 (4), 1363-5CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The form factor of a micelle model with a spherical core and Gaussian polymer chains attached to the surface was calcd. anal. and the results were compared to Monte Carlo simulations. Excluded vol. interactions between the core and the polymers were introduced in the simulations. The expansion of the coils due to this effect can be mimicked in the anal. calcns. by moving the center of mass of the chains away from the surface of the core. The anal. expression for the form factor has been used for analyzing small-angle scattering data.
- 67Balmer, J. A.; Mykhaylyk, O. O.; Schmid, A.; Armes, S. P.; Fairclough, J. P. A.; Ryan, A. J. Characterization of Polymer-Silica Nanocomposite Particles with Core-Shell Morphologies Using Monte Carlo Simulations and Small Angle X-Ray Scattering. Langmuir 2011, 27 (13), 8075– 8089, DOI: 10.1021/la201319h67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXntlCqtrg%253D&md5=9ea3ca1b4a8edcc48f631a3152519b91Characterization of Polymer-Silica Nanocomposite Particles with Core-Shell Morphologies using Monte Carlo Simulations and Small Angle X-ray ScatteringBalmer, Jennifer A.; Mykhaylyk, Oleksandr O.; Schmid, Andreas; Armes, Steven P.; Fairclough, J. Patrick A.; Ryan, Anthony J.Langmuir (2011), 27 (13), 8075-8089CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A two-population model based on std. small-angle x-ray scattering (SAXS) equations is verified for the anal. of core-shell structures comprising spherical colloidal particles with particulate shells. First, Monte Carlo simulations of core-shell structures are performed to demonstrate the applicability of the model. Three possible shell packings are considered: ordered silica shells due to either charge-dependent repulsive or size-dependent Lennard-Jones interactions or randomly arranged silica particles. In most cases, the two-population model produces an excellent fit to calcd. SAXS patterns for the simulated core-shell structures, together with a good correlation between the fitting parameters and structural parameters used for the simulation. The limits of application are discussed, and then, this two-population model is applied to the anal. of well-defined core-shell vinyl polymer/silica nanocomposite particles, where the shell comprises a monolayer of spherical silica nanoparticles. Comprehensive SAXS anal. of poly(styrene-co-Bu acrylate)/silica colloidal nanocomposite particles (prepd. by the in situ emulsion copolymn. of styrene and Bu acrylate in the presence of a glycerol-functionalized silica sol) allows the overall core-shell particle diam., the copolymer latex core diam. and polydispersity, the mean silica shell thickness, the mean silica diam. and polydispersity, the vol. fractions of the two components, the silica packing d., and the silica shell structure to be obtained. These exptl. SAXS results are consistent with electron microscopy, dynamic light scattering, TG, helium pycnometry, and BET surface area studies. The high electron d. contrast between the (co)polymer and the silica components, together with the relatively low polydispersity of these core-shell nanocomposite particles, makes SAXS ideally suited for the characterization of this system. Also, these results can be generalized for other types of core-shell colloidal particles.
- 68Matsen, M. W.; Thompson, R. B. Equilibrium Behavior of Asymmetric ABA Triblock Copolymer Melts. J. Chem. Phys. 1999, 111, 7139, DOI: 10.1063/1.48000668https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXmt1Oisr0%253D&md5=9e964439283738bcd88497f688856f2dEquilibrium behavior of symmetric ABA triblock copolymer meltsMatsen, M. W.; Thompson, R. B.Journal of Chemical Physics (1999), 111 (15), 7139-7146CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Melts of ABA triblock copolymer mols. with identical end blocks are examd. using SCF theory. Phase diagrams are calcd. and compared with those of homologous AB diblock copolymers formed by snipping the triblocks in half. This creates addnl. end segments which decreases the degree of segregation. Consequently, triblock melts remain ordered to higher temps. than their diblock counterparts. We also find that middle-block domains are easier to stretch than end-block domains. As a result, domain spacings are slightly larger, the complex phase regions are shifted towards smaller A-segment compns., and the perforated-lamellar phase becomes more metastable in triblock melts as compared to diblock melts. Although triblock and diblock melts exhibit very similar phase behavior, their mech. properties can differ substantially due to triblock copolymers that bridge between otherwise disconnected A domains. We evaluate the bridging fraction for lamellar, cylindrical, and spherical morphologies to be about 40%-45%, 60%-65%, and 75%-80%, resp. These fractions only depend weakly on the degree of segregation and the copolymer compn.
- 69Ryan, A. J.; Mai, S. M.; Fairclough, J. P. A.; Hamley, I. W.; Booth, C. Ordered Melts of Block Copolymers of Ethylene Oxide and 1,2-Butylene Oxide. Phys. Chem. Chem. Phys. 2001, 3 (15), 2961– 2971, DOI: 10.1039/b102784k69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlt1ygu74%253D&md5=be0263f79a0e3ef9cf50ea3c3e86b3abOrdered melts of block copolymers of ethylene oxide and 1,2-butylene oxideRyan, Anthony J.; Mai, Shao-Min; Fairclough, J. Patrick A.; Hamley, Ian W.; Booth, ColinPhysical Chemistry Chemical Physics (2001), 3 (15), 2961-2971CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)An account is presented of recent work on specially synthesized diblock, triblock and cyclic block copolymers of ethylene oxide and 1,2-butylene oxide. Simultaneous small-angle X-ray scattering and differential scanning calorimetry have been used to investigate the effects of block architecture on the stabilities and structures of microphase-sepd. melts. Stable lamellar, hexagonal, body-centered cubic and gyroid phases were detected. Phase diagrams are compared, one with another and with those predicted by the exact self-consistent mean-field theory, and center-block stretching in lamellar phases is confirmed.
- 70Mai, Y.; Eisenberg, A. Self-Assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41 (18), 5969, DOI: 10.1039/c2cs35115c70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1aqsbvL&md5=37964bf5011f9bc7a6aa42c4b612cf91Self-assembly of block copolymersMai, Yiyong; Eisenberg, AdiChemical Society Reviews (2012), 41 (18), 5969-5985CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Block copolymer (BCP) self-assembly has attracted considerable attention for many decades. The present tutorial review introduces the primary principles of BCP self-assembly in bulk and in soln., by describing expts., theories, accessible morphologies, etc.
- 71Lynd, N. A.; Meuler, A. J.; Hillmyer, M. A. Polydispersity and Block Copolymer Self-Assembly. Prog. Polym. Sci. 2008, 33 (9), 875– 893, DOI: 10.1016/j.progpolymsci.2008.07.00371https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXht1Klt77P&md5=30b678d94f8baaad00be1d8bec1759e0Polydispersity and block copolymer self-assemblyLynd, Nathaniel A.; Meuler, Adam J.; Hillmyer, Marc A.Progress in Polymer Science (2008), 33 (9), 875-893CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. Block copolymers consist of two or more chem. distinct polymers that are covalently bound. These materials self-assemble into fascinating mesostructures with features on the nanometer length scale and have been the subject of intense research interest for about four decades. These efforts have generally focused on model block copolymer systems where the mol. wt. distributions of all blocks are very narrow. Traditionally, many block copolymer systems have been prepd. by living anionic polymn. and thus usually exhibit narrow mol. wt. distributions in all blocks. Therefore, the assumption of monodisperse blocks that greatly simplifies theor. work is on solid exptl. ground. Prepn. of block copolymers with relatively broad mol. wt. distributions in one or more block has become increasingly common, however, as use of synthetic techniques such as controlled radical polymn. has proliferated. A review. Advances in these techniques have increased the no. of monomers readily incorporated into block copolymers and potentially will drive com. costs down. These polymn. strategies often, however, result in broader mol. wt. distributions than are typically obtained using living anionic, cationic, or metal-catalyzed techniques; understanding polydispersity effects should aid deployment of these block copolymers in advanced materials applications. This review describes both theor. and exptl. investigations of the effects of polydispersity on the melt-phase morphol. behavior of block copolymers. The summary includes research efforts focused on both continuous mol. wt. distributions and multicomponent blends. The review concludes with a summary and outlook on the potential utility of polydispersity as a tool to tune the morphol. behavior of block copolymers.
- 72Stieger, M.; Pedersen, J. S.; Lindner, P.; Richtering, W. Are Thermoresponsive Microgels Model Systems for Concentrated Colloidal Suspensions? A Rheology and Small-Angle Neutron Scattering Study. Langmuir 2004, 20 (17), 7283– 7292, DOI: 10.1021/la049518x72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXls1Kkt7w%253D&md5=16b72e479d1666a9e50320ada43e0381Are Thermoresponsive Microgels Model Systems for Concentrated Colloidal Suspensions? A Rheology and Small-Angle Neutron Scattering StudyStieger, Markus; Pedersen, Jan Skov; Lindner, Peter; Richtering, WalterLangmuir (2004), 20 (17), 7283-7292CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The structure of concd. temp.-sensitive poly(N-isopropylacrylamide) (PNiPAM) microgel suspensions has been investigated employing rheol. and small-angle neutron scattering (SANS). A previously described model expression for the particle form factor Pinho(q) is extended by a model hard sphere structure factor S(q), and the av. radial d. profiles φ(r) are calcd. from the amplitude of the form factor A(q) and the structure factor S(q). By this procedure, a direct real space description of the spatial ordering in the neighborhood of a single particle is obtained. The overall particle size and the correlation length ξ of the concn. fluctuations of the internal polymer network decrease with concn., revealing the increasing compression of the spheres. Thus, the particle form factor Pinho(q) of the swollen PNiPAM microgels depends on concn. The particle-particle interaction potential does not change significantly between 25 and 32 °C. Even approx. 1 K below the lower crit. soln. temp. (LCST), the exptl. scattering intensity distributions I(q)/c are described very well by the hard sphere structure factor when an equiv. hard sphere particle size RHS and vol. fraction ηHS are used. Microgels with different degrees of crosslinking and particle size resemble true hard sphere behavior up to effective vol. fractions of φeff < 0.35. At higher effective vol. fractions φeff > 0.35 strong deviations from true hard spheres are obsd. Interpenetration of the outer, less crosslinked regions of the soft spheres as well as particle compression occurred at higher concns. In agreement with this, the equil. colloidal phase behavior and rheol. also has some features of soft sphere systems. At temps. well above the LCST, the interaction potential becomes strongly attractive and the collapsed microgel spheres form aggregates consisting of flocculated particles without significant long-range order. Hence, an attractive interaction potential in concd. suspensions of PNiPAM microgels leads to distinctively different structures as compared to attractive hard sphere colloids. When the peculiar structural properties of the PNiPAM microgels are considered, they can be used as model systems in colloidal science.
- 73De Kruif, C. G.; Briels, W. J.; May, R. P.; Vrij, A. Hard-Sphere Colloidal Silica Dispersions. the Structure Factor Determined with SANS. Langmuir 1988, 4 (3), 668– 676, DOI: 10.1021/la00081a02973https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXitFSqs7s%253D&md5=1ac17a1929e4ab37b21bec0554d72a89Hard-sphere colloidal silica dispersions. The structure factor determined with SANSDe Kruif, C. G.; Briels, W. J.; May, R. P.; Vrij, A.Langmuir (1988), 4 (3), 668-76CODEN: LANGD5; ISSN:0743-7463.Small-angle neutron-scattering expts. were performed on the D11 instrument at Grenoble. The colloidal dispersions in cyclohexane studied contained silica particles sterically stabilized by octadecyl chains terminally grafted to the surface. From dil. dispersions the mean particle radius and the radius distribution were obtained. With these single-particle parameters the scattering behavior can be modeled with a hard-sphere interaction potential at all vol. fractions up to .vphi. = 0.4-0.5. Polydispersity effects on the representation of the structure factor were explicitly taken into account. These results confirm and amplify previous results on the modeling of a silica dispersion as a hard-sphere supramol. fluid.
- 74Pedersen, J. S.; Gerstenberg, M. C. Scattering Form Factor of Block Copolymer Micelles. Macromolecules 1996, 29 (4), 1363– 1365, DOI: 10.1021/ma951211574https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XltlWjsA%253D%253D&md5=e6f549fc773b455a50c0d6e157ca4981Scattering Form Factor of Block Copolymer MicellesPedersen, Jan Skov; Gerstenberg, Michael C.Macromolecules (1996), 29 (4), 1363-5CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)The form factor of a micelle model with a spherical core and Gaussian polymer chains attached to the surface was calcd. anal. and the results were compared to Monte Carlo simulations. Excluded vol. interactions between the core and the polymers were introduced in the simulations. The expansion of the coils due to this effect can be mimicked in the anal. calcns. by moving the center of mass of the chains away from the surface of the core. The anal. expression for the form factor has been used for analyzing small-angle scattering data.
- 75Lesage De La Haye, J.; Martin-Fabiani, I.; Schulz, M.; Keddie, J. L.; D’Agosto, F.; Lansalot, M. Hydrophilic MacroRAFT-Mediated Emulsion Polymerization: Synthesis of Latexes for Cross-Linked and Surfactant-Free Films. Macromolecules 2017, 50 (23), 9315– 9328, DOI: 10.1021/acs.macromol.7b0188575https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvVWrs7fL&md5=549d77c3d9ebacd5678b9afdaea0f5c7Hydrophilic MacroRAFT-Mediated Emulsion Polymerization: Synthesis of Latexes for Cross-Linked and Surfactant-Free FilmsLesage de la Haye, Jennifer; Martin-Fabiani, Ignacio; Schulz, Malin; Keddie, Joseph L.; D'Agosto, Franck; Lansalot, MurielMacromolecules (Washington, DC, United States) (2017), 50 (23), 9315-9328CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)A major drawback of conventional emulsion polymers arises from the presence of migrating low molar mass surfactants that contribute to poor water barrier properties and low adhesion to substrates. In this paper, we demonstrate how living polymer chains obtained by reversible addn.-fragmentation chain transfer (RAFT) can be used as an efficient stabilizer in emulsion polymn., leading to the prodn. of surfactant-free latexes, which then form crosslinked films with beneficial properties. Hydrophilic poly(methacrylic acid) (PMAA) chains obtained by RAFT performed in water are used to mediate emulsion polymn. and produce film-forming latex particles from mixts. of Me methacrylate, Bu acrylate, and styrene. Stable dispersions of particles with sizes between 100 and 200 nm are obtained, with very low amts. of coagulum (<0.5 wt. %). The particles are stabilized by the PMAA segment of amphiphilic block copolymers formed during the polymn. Remarkably, low amts. of PMAA chains (from 1.5 down to 0.75 wt. %) are enough to ensure particle stabilization. Only traces of residual PMAA macroRAFT agents are detected in the final latexes, showing that most of them are successfully chain extended and anchored on the particle surface. The glass transition temp. of the final material is adjusted by the compn. of the hydrophobic monomer mixt. so that film formation occurs at room temp. Conventional crosslinking strategies using addnl. hydrophobic comonomers, such as 1,3-butanediol diacrylate (BuDA), diacetone acrylamide (DAAm), and (2-acetoacetoxy)ethyl methacrylate (AAEM), are successfully applied to these formulations as attested by gel fractions of 100%. When particles are internally crosslinked with BuDA, chain interdiffusion between particles is restricted, and a weak and brittle film is formed. In contrast, when DAAm-contg. chains undergoes crosslinking during film formation, full coalescence is achieved along with the creation of a crosslinked network. The resulting film has a higher Young's modulus and tensile strength as a result of crosslinking. This synthetic strategy advantageously yields a surfactant-free latex that can be formed into a film at room temp. with mech. properties that can be tuned via the crosslinking d.
- 76Martín-Fabiani, I.; Lesage De La Haye, J.; Schulz, M.; Liu, Y.; Lee, M.; Duffy, B.; D’Agosto, F.; Lansalot, M.; Keddie, J. L. Enhanced Water Barrier Properties of Surfactant-Free Polymer Films Obtained by MacroRAFT-Mediated Emulsion Polymerization. ACS Appl. Mater. Interfaces 2018, 10 (13), 11221– 11232, DOI: 10.1021/acsami.8b0104076https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXktFSqu7s%253D&md5=eb3947a987c1093495342a2f1973200aEnhanced Water Barrier Properties of Surfactant-Free Polymer Films Obtained by MacroRAFT-Mediated Emulsion PolymerizationMartin-Fabiani, Ignacio; Lesage de la Haye, Jennifer; Schulz, Malin; Liu, Yang; Lee, Michelle; Duffy, Brendan; D'Agosto, Franck; Lansalot, Muriel; Keddie, Joseph L.ACS Applied Materials & Interfaces (2018), 10 (13), 11221-11232CODEN: AAMICK; ISSN:1944-8244. (American Chemical Society)The presence of low-molar-mass surfactants in latex films results in detrimental effects on their water permeability, gloss, and adhesion. For applications such as coatings, there is a need to develop formulations that do not contain surfactants and have better water barrier properties. Having previously reported the synthesis of surfactant-free latex particles in water using low amts. (<2 wt %) of chains synthesized by controlled radical polymn., here we study the water barrier properties of films made from these particles and their application in anticorrosion coatings. When films cast from aq. dispersions of acrylate copolymer particles stabilized with poly(sodium 4-styrenesulfonate) (PSSNa) were immersed in water for 3 days, they sorbed only 4 wt % water. This uptake is only slightly higher than the value predicted for the pure copolymer, indicating that the neg. effects of any particle boundaries and hydrophilic-stabilizing mols. are minimal. This sorption of liq. water is 5 times lower than what is found in films cast from particles stabilized with the same proportion of poly(methacrylic acid) (PMAA), which is more hydrophilic than PSSNa. In water vapor with 90% relative humidity, the PSSNa-based film had an equil. sorption of only 4 wt %. A small increase in the PMAA content has a strong and neg. impact on the barrier properties. NMR relaxometry on polymer films after immersion in water shows that water clusters have the smallest size in the films contg. PSSNa. Furthermore, these films retain their optical clarity during immersion in liq. water for up to 90 min, whereas all other compns. quickly develop opacity ("water whitening") as a result of light scattering from sorbed water. This implies a remarkably complete coalescence and a very small d. of defects, which yields properties matching those of some solvent-borne films. The latex stabilized with PSSNa is implemented as the binder in a paint formulation for application as an anticorrosive barrier coating on steel substrates and evaluated in accelerated weathering and corrosion tests. Our results demonstrate the potential of self-stabilized latex particles for the development of different applications, such as waterborne protective coatings and pressure-sensitive adhesives.
- 77Singh, K. B.; Tirumkudulu, M. S. Cracking in Drying Colloidal Films. Phys. Rev. Lett. 2007, 98 (21), 218302, DOI: 10.1103/PhysRevLett.98.21830277https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlvVagt78%253D&md5=ba2c452e115222337d1dee73f6c39e44Cracking in Drying Colloidal FilmsSingh, Karnail B.; Tirumkudulu, Mahesh S.Physical Review Letters (2007), 98 (21), 218302/1-218302/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)It has long been known that thick films of colloidal dispersions such as wet clays, paints, and coatings crack under drying. Although capillary stresses generated during drying have been recently identified as the cause for cracking, the existence of a max. crack-free film thickness that depends on particle size, rigidity, and packing has not been understood. Here, we identify two distinct regimes for crack-free films based on the magnitude of compressive strain at the max. attainable capillary pressure and show remarkable agreement of measurements with our theory. We anticipate our results to not only form the basis for design of coating formulations for the paints, coatings, and ceramics industry but also assist in the prodn. of crack-free photonic band gap crystals.
- 78Roberts, C. C.; Francis, L. F. Drying and Cracking of Soft Latex Coatings. J. Coatings Technol. Res. 2013, 10 (4), 441– 451, DOI: 10.1007/s11998-012-9425-778https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhtVGrtL3I&md5=3a2d5ac7b9d2c3a16b6d2e00f51d7215Drying and cracking of soft latex coatingsRoberts, Christine C.; Francis, Lorraine F.Journal of Coatings Technology and Research (2013), 10 (4), 441-451CODEN: JCTRCP; ISSN:1935-3804. (Springer)The min. film formation temp. (MFFT) is the min. drying temp. needed for a latex coating to coalesce into an optically clear, dense crack-free film. To better understand the interplay of forces near this crit. temp., cryogenic SEM (cryoSEM) was used to track the latex particle deformation and water migration in coatings dried at temps. just above and below the MFFT. Although the latex particles completely coalesced at both temps. by the end of the drying process, it was discovered that particle deformation during the early drying stages was drastically different. Below the MFFT, cracks initiated just as menisci began to recede into the packing of consolidated particles, whereas above the MFFT, partial particle deformation occurred before menisci entered the coating and cracks were not obsd. The spacing between cracks measured in coatings dried at varying temps. decreased with decreasing drying temp. near the MFFT, whereas it was independent of temp. below a crit. temp. Finally, the addn. of small amts. of silica aggregates was found to lessen the cracking of latex coatings near the MFFT without adversely affecting their optical clarity.
- 79Cowie, J. M. G.; Arrighi, V. Polymers: Chemistry and Physics of Modern Materials, 3rd ed.; CRC Press: Boca Raton, 2007.There is no corresponding record for this reference.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.macromol.2c01299.
Further characterization of triblock copolymer solutions, dispersions films, and SAXS models (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.