首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Z.L. Yao 《Polymer》2011,52(17):3769-3775
Well-defined statistical copolymer of poly (di(ethylene glycol) methyl ether methacrylate-stat-oligo(ethylene glycol) methyl ether methacrylate-C60 ((PMEO2MA-stat-POEGMA300)-C60) was synthesized via atom transfer radical polymerization (ATRP) reaction and atom transfer radical addition (ATRA) processes. The lower critical solution temperature (LCST) of PMEO2MA-stat-POEGMA300 increased from 42 to 95 °C when the amounts of methanol was increased from 0 to 30 vol%, beyond which the LCST could not be quantified. Similarly, the LCST of (PMEO2MA-stat-POEGMA300)-C60 also increased with methanol content, however it was lower than PMEO2MA-stat-POEGMA300 for all methanol/water compositions. The CMC of (PMEO2MA-stat-POEGMA300)-C60 increased with increasing methanol content, suggesting that methanol is a better solvent for PMEO2MA-stat-POEGMA300 segment. The amphiphilic (PMEO2MA-stat-POEGMA300)-C60 structure formed spherical micelles in water/methanol mixture, and larger micelles were formed at higher methanol content. The hydrodynamic radius (Rh) remained constant at temperature below the LCST. It increased dramatically at temperature greater than the LCST, and the (Rg/Rh) increased from ∼0.75 to ∼1.0. We believe that the (PMEO2MA-stat-POEGMA300) coronas dehydrate at higher temperature, and the micelles associate to form larger aggregates. In water/methanol mixtures, core-shell micelles and large compound micelles are produced below and above the LCST respectively.  相似文献   

2.
Xuezhi Tang 《Polymer》2007,48(21):6354-6365
Novel amphiphilic block copolymers, poly(ethylene oxide)-b-poly(p-nitrophenyl methacrylate) (PEO-b-PNPMA) with controlled molecular weights and narrow molecular weight distribution were successively synthesized by ATRP of NPMA using PEO-Br as initiator. Self-assembling of the diblock copolymer PEO113-b-PNPMA28 in the different solvent mixtures yielded various morphologies of star micelle-like aggregates, such as spheres, vesicles, cauliflower-like aggregates and rod-like aggregates, which are determined by the nature of the common solvents and the selective solvents. Thus the critical selective solvent contents and the solvent contents in PNPMA-rich phase were measured, and they have the following order: ethanol > methanol > water, and THF > CH3NO2 > DMSO. The probable self-assembling mechanism is discussed. This method is convenient for preparation of multiple morphological star micelle-like aggregates in solution, especially from the amphiphilic block copolymers with relatively longer block shell.  相似文献   

3.
A series of LiNi1/3Mn1/3Co1/3O2 samples with α-NaFeO2 structure belonging to the D3d5 space group were synthesized using tartaric acid as a chelating agent by wet-chemical method. Different acid to metal-ion ratios R have been used to investigate the effect of this parameter on the physical and electrochemical properties. We have characterized the reaction mechanism, the structure, and morphology of the powders by TGA, XRD, SEM and TEM imaging, completed by magnetic measurements, Raman scattering spectroscopy, and complex impedance experiments. We find that the LiNi1/3Mn1/3Co1/3O2 sintered at 900 °C for 15 h with an acid to metal-ion ratio R = 2 was the optimum condition for this synthesis. For this optimized sample, only 1.3% of nickel-ions occupied the 3b Wyckoff site of the lithium-ions sublattice. The electrochemical performance has been investigated using a coin-type cell containing Li metal as the anode. The electronic performance is correlated to the concentration of the Ni(3b) defects that increase the charge transfer resistance and reduce the lithium diffusion coefficient. The optimized cell delivered an initial discharge capacity of 172 mAh g−1 in the cut-off voltage of 2.8-4.4 V, with a coulombic efficiency of 93.4%.  相似文献   

4.
B.W. Mao  Y.Y. Gan  O.K. Tan 《Polymer》2005,46(23):10045-10055
Diblock copolymers of t-butyl methacrylate (tBMA) and 2-(diethylamino)ethyl methacrylate (DEAEMA) were successfully synthesized by one-pot strategy via the atom transfer radical polymerization (ATRP). Kinetic results clearly demonstrated the controlled/‘living’ character of the polymerization. The zwitterionic block copolymers of poly(methacrylic acid-b-DEAEMA), obtained by hydrolysis of poly(tBMA-b-DEAEMA), showed pH-dependent reverse micellization behavior. Micellar aggregates formed from poly(MAA30-b-DEAEMA71), poly(MAA68-b-DEAEMA55) and poly(MAA64-b-DEAEMA44) had fairly low polydispersity index at both solutions of low pH of 2 and high pH of 12. Micelles formed at pH 2 were larger (Rh∼40-61 nm) with looser core due to hydration of the MAA. In the presence of simple electrolyte (0.3 mol dm−3 NaCl solution), the size of the micelles reduced by almost half while the aggregation number was little changed. This is attributed to the draining of the hydrated micellar core due to osmotic pressure. On the other hand, DEAEMA-core micelles formed at pH 12 were compact and much smaller (Rh∼14-22 nm). Addition of NaCl had only a small effect. The micellar size reduced only slightly due to the electrostatic screening effect and the aggregation number was almost unchanged.  相似文献   

5.
This paper reports on the microstructure of anodic titanium oxide (TiO2) and its use in a dye-sensitized solar cell (DSSC) device. When voltages of 60 V were applied to titanium foil for 2 hr under 0.25 wt% NH4F+ 2 vol% H2O+C2H4(OH)2, TiO2 with a nanotube structure was formed. The film, which had a large surface area, was used as an electron transport film in the DSSC. The DSSC device had a short-circuit current density (Jsc) of 12.52 mA cm−2, a fill factor (FF) of 0.65, an open-voltage (Voc) of 0.77 V, and a photocurrent efficiency of 6.3% under 100% AM 1.5 light. The internal impedance values under 100%, 64%, 11%, and 0% (dark) AM 1.5 light intensities were measured and simulated using the electrical impedance spectroscopy (EIS) technique. The impedance characteristics of the DSSC device were simulated using inductors, resistors, and capacitors. The Ti/TiO2, TiO2/Electrolyte, electrolyte, and electrolyte/(Pt/ITO) interfaces were simulated using an RC parallel circuit, and the bulk materials, such as the Ti, ITO and conducting wire, were simulated using a series of resistors and inductors. The impedance of the bulk materials was simulated using L0+R0+Rb, the impedance of the working electrode was simulated using (C1//R1)//(Ra+(C2//R2), the electrolyte was simulated using C3//R3, and the counter electrode was simulated using C4//R4.  相似文献   

6.
The dimensions of linear polymer chains are scaled to their molar mass (M) as R = kMα with α = 1/2 and 3/5 in a theta and an athermal solvent, respectively. In a good solvent, both k and α are a function of the solvent quality and chain length range. A high-temperature laser light-scattering spectrometer was used to measure the average radius of gyration (〈Rg〉) and hydrodynamic radius (〈Rh〉) of a set of narrowly distributed linear polystyrene chains in decalin over a wide temperature range. k and α in the scaling experimentally varying with T over a chain length range was analyzed. The results reveal that for 〈Rg〉, α = 0.59 − 0.09exp(−τ/0.066) and k = 0.60τ2α−1, reasonably agreeing with the thermal blob theory. For 〈Rh〉, α = 0.59 − 0.09exp(−τ/0.106), but k deviates from the relationship of k ∝ τ2α−1, reflecting that the hydrodynamic interaction and chain draining are not considered in the thermal blob theory.  相似文献   

7.
P. Ravi  L.H. Gan  Y.Y. Gan  X.L. Xia  X. Hu 《Polymer》2005,46(1):137-146
Homopolymers of azobenzene (azo) methacrylates with different substituents and their diblock copolymers with poly(2-(dimethylamino)ethyl methacrylate p(DMAEMA) were synthesized via atom transfer radical polymerization (ATRP). Controlled/‘living’ ATRP of azo methacrylates were achieved up to ∼50% conversion, after which deviation occurred. It was found that the copolymerization rate of 6-[4-phenylazo]phenoxy]hexylmethacrylate (PPHM) from p(DMAEMA) macroinitiator was almost identical to that for the homopolymerization of PPHM monomer, with kapp∼0.0078 min−1. For the copolymerizations, almost complete incorporation of the azo methacrylate monomers could be obtained with low molecular weight macroinitiator (PDMAEMA)-Cl, whereas macroinitiators of long chain length did not give full conversion, most likely due to chain floding and steric hindrance caused by the bulk azo monomers. Because azo monomers are highly hydrophobic, only the diblock copolymers with short azo segment were soluble in water which self-assembled into micellar particles. The effect of photo-induced trans-cis isomerization on lower critical solution temperature (LCST) and surface tension were studied. The LCST of the diblock copolymers increased upon irradiation by UV light due to the cis conformers being more hydrophilic. However, the trans-cis isomerization had only small effect on the critical micelle concentration (cmc) and γcmc of azo methacrylate block copolymers, due to the formation of compact core of the micelles. The formations of core-shell micelles were established from LLS and TEM studies. All the three azo methacrylate amphiphilic block copolymers formed hard core-shell micelles with relatively small Rh values of 31 nm for p(DMAEMA172-b-BPHM7), 26 nm for p(DMAEMA172-b-CPHM7) and 32 nm for p(DMAEMA172-b-PPHM9). Whereas for the azo acrylate copolymer, p(DMAEMA172-b-BPHA6), large micelles with Rh∼78 nm with loose core was formed.  相似文献   

8.
Soo-Young Park  Woo-Hwan Sul 《Polymer》2008,49(15):3327-3334
The effects of the solvent selectivity of toluene/ethanol mixtures on the micellar and ordered structures of an asymmetric diblock copolymer of PS(19.6 K)-b-P4VP(5.1 K) in the dilute (1 wt%) and semi-dilute (8 wt%) solutions, as well as in the gel and solid films, were studied using small angle X-ray scattering (SAXS), generalized indirect Fourier transform (GIFT), and transmission electron microscopy (TEM) methods. The solvent selectivity was controlled by ? (weight percentage of ethanol in toluene/ethanol mixture). Individual micelles, space-filled micellar structure (without three-dimensional order), and three-dimensionally ordered gel and solid structures were observed from the 1 and 8 wt% solutions, the gel, and the solid film, respectively. In the 1 wt% solution, the individual micellar structures were strongly dependent on ?; the spherical micelles with P4VP core at ? = 0, the unimer state at 10 ≤ ? ≤ 50, the spherical micelles with PS core at ? = 60, the cylindrical micelles with PS core at ? = 70 and 80, and precipitation at ? = 90 and 100 were observed. The 8 wt% solution was close to overlap concentration with the unimer state in the regions of 20 ≤ ? ≤ 40. In the gel, the ordered structure was observed in the sequence of bcc, hexagonal, gyroid, lamellar, reverse hexagonal and random as ? increased, and could be explained by the change of the relative volume fraction of each block as ? changed, similar to the phase sequence in the phase diagram of the diblock copolymer. The solid films showed the various kinetically frozen ordered microstructures such as randomly packed sphere, hexagonal, gyroid, hexagonally perforated lamella, reversed hexagonal, and randomly packed cylinder, which were controlled by the solvent quality in the gel before solidification. We believe that these results can be applied to photonic crystals, self-assembled nano-patterning, and functional nanoparticles in which the structural control is most important.  相似文献   

9.
LiNi1−xCoxO2 (x = 0, 0.1, 0.2) cathode materials were successfully synthesized by a rheological phase reaction method with calcination time of 0.5 h at 800 °C. All obtained powders are pure phase with α-NaFeO2 structure (R-3m space group). The samples deliver an initial discharge capacity of 182, 199 and 189 mAh g−1 (25 mA g−1, 4.35-3.0 V), respectively. The reaction mechanism was also discussed, which consists of a series of defect reactions. As a result of these defect reactions, the reaction of forming LiNi1−xCoxO2 takes place in high speed.  相似文献   

10.
A series of poly[2-(diisopropylamino)ethyl methacrylate]-block-poly[2-(N-morpholino)ethyl methacrylate], [PDPA-b-PMEMA], have been synthesized by using group transfer polymerization. These novel PDPA-b-PMEMA diblock copolymers dissolved molecularly in aqueous solution at low pH (<6.0) due to the protonation of all tertiary amine residues of both blocks and formed PDPA-core micelles at pH 7.5 by PMEMA block forming the micelle coronas. On the other hand, it was also observed that these diblock copolymers formed near-monodisperse ‘reverse micelles’, PMEMA-core micelles, in n-alkanes with or without requiring cosolvent depending on comonomer ratios. Dynamic light scattering studies indicated monodisperse or near-monodisperse micelles in both cases. The intensity-average radii of the PDPA-core and the PMEMA-core micelles were between 10 nm and 17 nm (polydispersity index, μ2/Γ2 < 0.08) and between 10 nm and 13 nm in n-hexane (μ2/Γ2 < 0.09), respectively.  相似文献   

11.
Xiongxiong Luo 《Polymer》2008,49(16):3457-3461
Controlled/living radical polymerization of styrene has been achieved by atom transfer radical polymerization (ATRP) catalyzed by cobaltocene (PDI = 1.27-1.41). The effects of the initiators, temperatures and solvents were studied. The end group of PS-Br was characterized by 1H NMR. Block copolymerization proved that the polymer end is still living and the PMMA-b-PSt block copolymer was synthesized.  相似文献   

12.
This paper reports the synthesis and magnetism of a new polymer-inorganic intercalation nanocomposite based on a C60-containing poly(ethylene oxide) (C60-PEO) into layered MnPS3, which is characterized by XRD, IR and thermal analyses. The lattice expansion (Δd) of the intercalation nanocomposite is about 9.3 Å indicating the successful intercalation. And the charge balance is maintained by K+ ions coordinating with PEO chain of C60-PEO polymer, which come from the pre-intercalation compound Mn1−xPS3[K2x(H2O)y]. Magnetic measurements indicate that the intercalation nanocomposite (C60-PEO/MnPS3) exhibits a magnetic phase transition from paramagnetism to ferrimagnetism at about 40 K. And the distinctive hysteresis of M-H relationship further confirms that it is a low temperature ferrimagnetic nanocomposite.  相似文献   

13.
Surface-functionalized polymeric nanoparticles were prepared by: a) self-assembly of poly(4-vinylbenzocyclobutene-b-butadiene) diblock copolymer (PVBCB-b-PB) to form spherical micelles (diameter: 15-48 nm) in decane, a selective solvent for PB, b) crosslinking of the PVBCB core through thermal dimerization at 200-240 °C, and c) cleavage of the PB corona via ozonolysis and addition of dimethyl sulfide to afford aldehyde-functionalized nanoparticles (diameter: ∼16-20 nm), along with agglomerated nanoparticles ranging from ∼30 to ∼100 nm in diameter. The characterization of the diblock copolymer precursors, the intermediate micelles and the final surface-functionalized crosslinked nanoparticles was carried out by a combination of size exclusion chromatography, static and dynamic light scattering, viscometry, thermogravimetric analysis, 1H NMR and FTIR spectroscopy and transmission electron microscopy.  相似文献   

14.
Surface-enhanced Raman scattering (SERS) was used to investigate C60 self-assembling in solvents like pyrrolidine (Py) and N-methyl-2-pyrrolidinone (NMP) as well as in binary mixtures of o-dichlorobenzene (DCB)/acetonitrile (ACN) and DCB/NMP. For a correct evaluation of the modifications of Raman spectra induced by the C60 aggregation, we have also presented the variations due to the measuring method, i.e., the signal dependence of the metallic support type and the surface roughness. The interaction between C60 and the Au substrate, appearing as a chemical component in SERS generation, is mainly evidenced by a band at ∼342 cm−1. In the aggregated phase, the intermolecular interactions lead to a reduction in the parent Ih C60 symmetry as observed by a modified phonon spectrum. As a general feature, the spectral range below 800 cm−1 is the most diagnostic for the aggregate assignment, the main indicative being the disappearance of the Raman bands associated to the radial vibration modes. SERS measurements have revealed two stages in the self-assembling of C60 in NMP. In the beginning, charge-transfer molecular complexes that associate slowly in stable aggregates are formed by the binding of an NMP molecule to the C60 cage. These complexes are noticed in the SERS spectrum by the replacement of the original Hg(1) band at ∼269 cm−1 with two others at ∼255 and ∼246 cm−1. In the aggregated phase, when using NMP and P as a solvent, the Raman spectrum reveals new bands that appear around 94 and 110-118 cm−1, which are associated with the interball interactions. In a DCB/ACN solvent mixture, the self-assembling process is driven by weak van der Waals type forces and resembles a precipitation, yielding C60 clusters of different size.  相似文献   

15.
Pure and mixed gas n-C4H10 and CH4 sorption and dilation in poly(1-trimethylsilyl-1-propyne) (PTMSP) are reported at temperatures ranging from −20 to 35 °C. The presence of n-C4H10 in the mixture considerably reduces CH4 solubility. For example, CH4 solubility (in the limit of zero CH4 fugacity) at 25°C decreases from 4.0 (pure gas) to 0.78 cm3(STP)/(cm3 polymer atm) in the presence of n-C4H10 at an activity of 0.60. At −20 °C, CH4 solubility decreases by almost an order of magnitude, from 10.2 (pure gas) to 1.22 cm3(STP)/(cm3 polymer atm) in the presence of n-C4H10 at an activity of 0.61. In contrast, n-C4H10 mixture sorption properties are not measurably affected by the presence of CH4. The dual mode sorption model parameters for CH4 and n-C4H10 in PTMSP were determined from pure and mixed gas sorption measurements, and this model can adequately describe the sorption data. The n-C4H10/CH4 mixed gas solubility selectivity in PTMSP decreases as temperature increases and as n-C4H10 activity increases. For example, at 25 °C, the n-C4H10/CH4 solubility selectivity decreases from 250 to 120 as n-C4H10 activity increases from 0.02 to 0.25. At −20 °C and an n-C4H10 activity of 0.24, the n-C4H10/CH4 solubility selectivity is 590. Penetrant-induced volume dilation of PTMSP can be adequately modeled by assuming that all swelling is caused by penetrant molecules sorbed in the polymer's dense equilibrium region (i.e., the Henry's law region) during sorption. However, the best fit partial molar volumes in the Henry's law region for the dilation data are considerably lower than the penetrant partial molar volumes in liquids, suggesting that further theoretical efforts are needed to develop predictive models of volume dilation in high free volume glassy polymers.  相似文献   

16.
Layered metastable lithium manganese oxides, Li2/3[Ni1/3−xMn2/3−yMx+y]O2 (x = y = 1/36 for M = Al, Co, and Fe and x = 2/36, y = 0 for M = Mg) were prepared by the ion exchange of Li for Na in P2-Na2/3[Ni1/3−xMn2/3−yMx+y]O2 precursors. The Al and Co doping produced the T#2 structure with the space group Cmca. On the other hand, the Fe and Mg doped samples had the O6 structure with space group R-3m. Electron diffraction revealed the 1:2 type ordering within the Ni1/3−xMn2/3−yMx+yO2 slab. It was found that the stacking sequence and electrochemical performance of the Li cells containing T#2-Li2/3[Ni1/3Mn2/3]O2 were affected by the doping with small amounts of Al, Co, Fe, and Mg. The discharge capacity of the Al doped sample was around 200 mAh g−1 in the voltage range between 2.0 and 4.7 V at the current density of 14.4 mA g−1 along with a good capacity retention. Moreover, for the Al and Co doped and undoped oxides, the irreversible phase transition of the T#2 into the O2 structure was observed during the initial lithium deintercalation.  相似文献   

17.
Uma Chatterjee 《Polymer》2005,46(5):1575-1582
ATRP of several methacrylates viz. methyl methacrylate (MMA), ethyl methacrylate (EMA), n-butyl methacrylate (nBMA), t-butyl methacrylate (tBMA), benzyl methacrylate (BzMA) and (N,N-dimethylamino)ethyl methacrylate (DMAEMA) has been studied in neat as well as aqueous (up to 12 vol% water) acetone at 35 °C using CuCl/bipyridine (bpy) catalyst and ethyl 2-bromoisobutyrate as the initiator. Addition of water significantly enhances the rate of polymerization without losing control. Unlike CuCl/bpy the CuBr/bpy catalyst gives poor control which is attributed to the lower solubility and consequent heterogeneity in the latter case. Of the other ligands used with the CuCl catalyst viz. o-phenanthroline (o-phen), 1,1,4,7,7-pentamethyldiethylenetriamine (PMDETA), 1,1,4,7,10,10-hexamethyltriethylenetetramine (HMTETA), Me6TREN only o-phen offers reasonably good control. The CuCl/bpy catalyst system has been used also in preparing some di- and tri-block copolymers with reasonably low polydispersity index (PDI) at ambient temperature (35 °C) using aqueous acetone as the solvent. The following block copolymers have been prepared PMMA-tBMA, PMMA-b-tBMA-b-MMA, PMMA-DMAEMA, by this method.  相似文献   

18.
Fullerene-activated carbon composite electrodes were prepared and their charge/discharge characteristics were studied for use in a high power electric double-layer capacitor. The capacitance of the C60-loaded activated carbon fiber (ACF) electrodes became greater than that of the unloaded ACF at charge/discharge current densities above 50 mA/cm2. In order to obtain a highly dispersed C60-loaded electrode, an ultrasonic treatment was performed. The size of the C60 agglomerate decreased from 1-2 to 0.1 μm or less, and the capacitance of the C60-loaded ACF electrodes increased with an increase in the ultrasonic treatment time. A higher capacitance of 172 F/g was obtained at 50 mA/cm2 on a 1 wt% C60-loaded electrode with ultrasonic treatment, and the C60-loaded ACF electrode also showed a higher cycle performance.  相似文献   

19.
At room temperature, the hexagonal C60. 2(CH3)CCl3 solvate (a = 10.13(1) Å, c = 10.84(1) Å), made of alternating layers of C60 and solvent molecules, forms with a negative excess volume, and its desolvation enthalpy is virtually the same as the sublimation enthalpy of the pure solvent. Crystallographic and calorimetric studies vs temperature indicate that hexagonal C60. 2(CH3)CCl3 changes at 211.7 K (1.3 kJ mol−1 of solvate) into an intermediate triclinic phase which transforms at 189.7 K (4.1 kJ mol−1 of solvate) into another triclinic phase.A crystallographic analysis in the series of hexagonal C60. 2 YCCl3 solvates (Y = H, Cl, Br, CH3) reveals that: (i) the change in the unit-cell volume values is due to a change in axis c whose value depends on the size of Y, (ii) the molar volume of the solvates depends linearly on the molar volume of the solvents.Ageing studies at room temperature show that C60. 2(CH3)CCl3 loses its solvent molecules within a few days or a few months, depending on storage conditions.  相似文献   

20.
The dynamics of phase separated micellar solutions of randomly sulphonated polystyrene ionomers in toluene were examined with neutron spin echo spectroscopy (NSE). The correlation functions demonstrated single exponential decays with a constant background offset. The relaxation rate (Γ) of the ionomer micelles at small length scales in the q range 0.075-0.16 Å−1 was diffusive (Γ ∼ q2) as expected for the collective breathing mode of a cross-linked gel. At intermediate length scales in the q-range 0.025-0.075 Å−1 the relaxation rates were non-diffusive (Γ ∼ q0.36, q0.72), which is attributed to the hopping dynamics of the individual stickers inside the ionomer micelles (τsticker ∼ 10 ns). At large length scales the scattering due to the phase separated inhomogeneities of the micellar network did not relax on the time scales of the measurements (<20 ns), giving rise to a constant background on the correlation functions. This slow relaxation process may be due to the hopping dynamics of whole micelles previously observed in rheology experiments (τmicelle ∼ 0.05 s). The NSE results are in agreement with a model developed in previous small-angle neutron scattering and rheology experiments for concentrated solution of ionomeric micelles. The NSE results for the associating ionomers are markedly different from the Zimm dynamics (Γ ∼ q3) previously observed for semi-dilute and cross-linked polystyrene polymers in a good solvent. The ionomeric cross-links thus have a large impact on the polymer chain dynamics at the nanosecond time scale.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号