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 78 毫秒
1.
Water-soluble copolymers, poly(dodecyl acrylate–coacrylic acid) (PDA), are synthesized by copolymerizing different ratios of dodecyl–acrylate/acrylic acid. Their aqueous solutions are transparent. The plots of I1/I3 (the intensity ratio of pyrene emission at 374 nm and 385 nm) and of surface tension as a function of polymer concentrations are combined to verify the existence of the aggregates of PDAs. It is found that the aggregations begin to form at concentrations below that of the polymer transfer to the air–water interface. The plots of I1/I3 as a function of polymer concentration shows that the polymers with a higher ratio of dodecyl acrylate form polymer aggregates at lower concentrations. The hydrophobicity of the inner core of polymer aggregates is close to that of sodium dodecyl sulfate (SDS). Polymer solutions of PDA 30, PDA 30L, and PDA 15, show significant solubilization ability to pyrene. (The molar ratio of dodecyl acrylate/total acrylates is 30%, 30%, and 15% for PDA 30, PDA 30L, and PDA 15, respectively. L, low molecular weight.) The solubilization ability increases with the increase in polymer concentration and degree of substitution of dodecylamine. The intensity ratio of the excimer emission to monomer emission of pyrene (Ie/Im) increases with increasing polymer concentration, roughly parallel to the increase in the amount of pyrene solubilized. However, the Ie/Im at a fixed pyrene concentration (Ie/I) decreases with increasing polymer concentration. These phenomena are interpreted in terms of the amount of pyrene solubilized and the size and concentration of the polymer aggregates. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Summary Possibility of living metathesis polymerization by Mo catalysts was examined for (p-n-butyl-o,o,m,m-tetrafluorophenyl)acetylene, which has two fluorine atoms at both ortho positions. The MoOCl4-n-Bu4Sn-EtOH (1:1:1) catalyst yielded a polymer with narrow molecular weight distribution ( ), but the corresponding MoCl5-based catalyst did not formed such a polymer. With the former catalyst, the number-average molecular weight of polymer increased in direct proportion to monomer conversion, while the molecular weight distribution remained narrow; this proves the livingness of the polymerization. The optimal conditions for the living polymerization were [n-Bu4Sn]/[MoOCl4]=∼1.0, [EtOH]/[MoOCl4]=0.5–1.5, and temperature≤30 °C. n-Butyl acetate and acetone as well as EtOH were effective as third catalyst components.  相似文献   

3.
The characterization of ethylene polymerization behaviors catalyzed over Cp2ZrCl2/MAO homogeneous system using methylaluminoxanes prepared by the direct hydrolysis of AlMe3 (Me=methy1) were reported. The MAO was prepared at the ratio of [H2O]/[A1]=1 and 0.5 and at three different temperatures, i.e., −40, −60 and −80 °C. The polymerization rate was not decreased with polymerization time when the MAO prepared at the ratio of [H2O]/[AlMe3]=l at −60 °C was used as a cocatalyst regardless of the ratio of Al/Zr and the polymerization temperature. The polymerization rate drastically decreased with polymerization time above 60 °C in case of using MAO prepared at the ratio of [H2O]/[AlMe3]=l at −80 °C. However, in case of the MAO prepared at the ratio of [H2O]/ [AlMe3]=0.5 at −80 °C, the rate continuously increased with polymerization time at the polymerization temperature of 70 °C and 80 °C. The amount of MAO needed to activate Cp2ZrC12 was larger than that of MAO prepared at the ratio of [H2O]/[A1]=1. The viscosity molecular weight of polyethylene (PE) cocatalyzed with MAO prepared at the ratio of [H2O]/[Al]=0.5 was lower than that of polyethylene obtained with MAO prepared at the ratio of [H2O]/[A1]=1.  相似文献   

4.
Both pyrene-fluorescence probe and fluorescence label techniques are used to investigate the association behaviors of hyperbranched poly(sulfone-amine) (HPSA) in aqueous solution. In the presence of HPSA, excimer emission peak evidently appeared, while no excimer peak was observed in the emission spectra in the absence of HPSA. The excitation spectrum monitored at excimer emission red shifts by about 38-40 nm compared to that monitored at monomer emission, which shows that the excimer is formed by preassociated pyrene chromophores. In the same concentration of pyrene, monomer emission of pyrene decreases but excimer emission increases with increasing the concentration of HPSA; the ratio of excimer-to-monomer emission intensity (IE/IM) gradually increases, reaches a critical point at 5-7 g/l, and sharply increases with the concentration. Pyrene-labeled hyperbranched poly(sulfone-amine) (Py-HPSA) was synthesized from 4-(1-Pyrene)butyroyl chloride and HPSA. The monomer emission and excimer emission of Py-HPSA show the concentration-quenching effect, while IE/IM increases monotonously, approaches a critical point, and then suddenly increases with increasing the concentration of Py-HPSA. Influences of acidity and solvents on the fluorescence emission were studied. In high concentrations of hyperbranched polymer, pH and DMSO significantly influence the emission of pyrene, and excimer peak disappears at 72% of DMSO fraction.  相似文献   

5.
The complexation between poly(ethylene oxide) (PEO) and poly(acrylic acid) (PAA) was made by using double the molar quantity of either polymer component at pH 2 where the resulting complex completely precipitates. After the removal of the precipitate, PEO or PAA remaining in the supernatant was subjected to gel permeation chromatography to investigate the change in the molecular weight distribution (MWD) caused by the complexation. No remarkable difference is observed in the MWD curves for PEO[1] (Mw=1.37 × 104) before and after the complexation with PAA[1] (Mw=1.10 × 103) and PAA[2] (Mw=4.16 × 105). However, the MWD curves of PEO[2] (Mw=1.26 × 105) and PAA[2] become shortened and shift to the low molecular weight side after the complexation with PAA[1] or [2] and PEO[2], respectively. This tendency is enhanced by increasing the complexation temperature. From these results, it is indicated that the complexation between PEO and PAA deals with an equilibrium reaction, and the equilibrium constant is dependent on the chain length of both polymer components and also on the complexation temperature.  相似文献   

6.
The kinetics of the ligand exchange reaction of the Cu(II)-ammine complex with poly(vinyl alcohol) (PVA) has been studied by a stopped-flow method at pH 9–10, at μ=0.1 (NH4Cl) and at 25°C. The reaction is initiated by the formation of unstable [Cu(NH3)3]2+ by the attack of H+ on Cu(II)-ammine complex, and proceeds through the mixed complex {[Cu(NH3)3(O?PVA)]2+}. This step may be rate-determining, followed by a rapid reaction. Finally, the Cu(II) ion is taken up by PVA. The rate is given by d[Cu(II)?PVA]/dt=k[H+]{[Cu(NH3)4]2+}[PVA]/[NH4Cl], where k=k1 + k2[H+], k1=4.25× 10s?1 and k2=5.20× 1011l mol?1s?1.  相似文献   

7.
Two asymmetric dithioether ligands with cyclohexyl (L1) and phenyl (L2) end-groups were synthesized. Reaction of L1 with copper(I) iodide afforded a 1D channel-type coordination polymer {[Cu4I4(L1)2](CH3CN)0.5}n (1) interconnected by cubane-type tetranuclear Cu4I4 cluster units. Whereas, a 2D brick-wall-type coordination polymer [CuI(L2)]n (2) with rhomboid dinuclear Cu–I2–Cu nodes was isolated from the reaction of L2 with copper(I) iodide.  相似文献   

8.
A water-soluble styrene–maleic anydride copolymer (SMA) is derivatized with different lipophilic groups, butylamine and dodecylamine, with different degrees of substitution (5, 15, and 30%). These lipophile-grafted SMAs are water-soluble, and their solutions are transparent. A plot of I1/I3 as a function of polymer concentrations indicates the extent of aggregate formation. Surface tension methods also verify the existence of aggregates. It is found that the aggregates begin to form at concentrations below that of the polymer transfer to the air–water interface. The plots of I1/I3 as a function of polymer concentrations for SMAs of different molecular weights derivatized with different lipophile with varying degrees of substitution show that the polymers with a higher degree of substitution and lower molecular weights aggregate at lower concentrations. Polymers substituted by butylamine form aggregates at a very high concentration, independent of the degree of sub-stitution. These phenomena are interpreted in terms of hydrophobic interactions as in micelles formed from surfactants. The higher degree of dodecyl-substituted SMA solubilizes pyrene at higher concentrations, and these pyrene solubilized solutions show pyrene excimer emission spectra. These emission spectra are used to characterize the relative size and hydrophobicity of aggregates. © 1993 John Wiley & Sons, Inc.  相似文献   

9.
A novel amino‐functionalized polystyrene copolymer (PS‐NH2) was designed and synthesized with styrene and 4‐vinylbenzyl amine. Additionally, an amino modified glass (G‐NH2) was obtained as a carrier. (PS‐NH2/pyrene)/G‐NH2 fluorescent nanofibrous membrane [named (PS‐NH2/pyrene)/G‐NH2] was designed and prepared via electrospinning technique to detect representative saturated nitroaromatic (NAC) explosive vapor. The (PS‐NH2/pyrene)/G‐NH2 showed highly fluorescence stability in ambient condition and further displayed a high quenching efficiency of 70.9% toward trinitrotoluene (TNT) vapor (~10 ppb) with an exposure time of 150 s at room temperature. The abundance of amino groups could effectively adsorb NACs and the binding of electron‐deficient NACs to the amino groups on the (PS‐NH2/pyrene)/G‐NH2 surface led to the formation of charge‐transfer complexes. The quenching constant (KSV) to TNT was obtained to be 1.07 × 1011 mL/g in gaseous phase with a limit of detection up to 2.76 × 10?13g/mL. Importantly, the (PS‐NH2/pyrene)/G‐NH2 showed notable selectivity toward TNT and 2,4‐dinitrotoluene vapors. Straightforwardly, the colorimetric sensing performance can be visualized by naked eye with a color change for detecting of different vapor phase NACs explosives. © 2018 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2018 , 135, 46708  相似文献   

10.
Equation of state for polymer solution   总被引:1,自引:0,他引:1  
S. Matsuoka  M.K. Cowman 《Polymer》2002,43(12):3447-3453
The flow pattern through a cloud of polymer segments is obviously different from the flow pattern around a solid object. It can be shown theoretically, however, that the partial viscosity due to the cloud can take the same value as for a solid sphere with the radius of gyration of the cloud as its radius. The specific viscosity of polymer solution has been derived as 2.5(c/cI), with cI being the internal concentration associated with a polymer molecule. The internal concentration is the ratio of mass over the volume of gyration of segments in a polymer chain. A radius of gyration exists for any type of polymers, flexible or rigid, exhibiting different kinds of dependence on the molecular weight. From the expression of the specific viscosity, the intrinsic viscosity is shown to be equal to 2.5/c, c being the (minimum) internal concentration for the state of maximum conformational entropy. The equation for the specific viscosity, thus obtained, is expanded into a polynomial in c[η]. This formula is shown to agree with data for several kinds of polymers, with flexible, semi-rigid and rigid.The quantity 1/cI can be interpreted as an expression for the chain stiffness. In polyelectrolytes, coulombic repulsive potentials affect the chain stiffness. The dependence of cI on the effective population of polyions in the polyelectrolyte molecule is discussed.An equation of state for the polymer solution is formulated that included the internal concentration. The virial coefficients emerge as a result of cI not always being equal to c, and they are molecular weight dependent.  相似文献   

11.
An in situ steady-state fluorescence (SSF) technique was applied in order to study the dissolution process of polystyrene (PS) latex films. The effect of the molecular weight M w of the PS on the dissolution rate was investigated. The PS chains were copolymerized with (1-pyrene)methyl methacrylate in order to make use of pyrene (P) as a fluorescent probe to monitor the dissolution process. Seven different films were prepared from P-labeled PS latex dispersions with different molecular weights at room temperature. These films were then annealed at 200 °C for 15 min to complete the film formation process before dissolution. The dissolution of PS films in a toluene (70 %)–cyclohexane (30 %) mixture was monitored in real time by watching the change in the fluorescence intensity of P, I P. We used a model that included both case I and case II diffusion kinetics to interpret the results of the dissolution experiments. The relaxation constants k 0 and the dissolution coefficients D d of the polymer chains were measured. Two different dissolution coefficients were obtained, which were attributed to the small and long polymer chains in the film, considering the high polydispersity of the polymer. It was also found that both of the D d values scaled with M w according to the law D d M ?n .  相似文献   

12.
We have synthesized a series of pyrene/cholesterol co-functionalized adlayers on quartz, oxidized silicon, indium-doped tin oxide and gold substrates. The pyrene derivative is N-1-pyrenesulfonyl-ethylenediamine (PSEDA) and the cholesterol derivative is cholesterol-ethylenediamine (Chol-NH2), which was bound covalently to substrates through epoxide functionalities. X-ray photoelectron spectroscopy shows covalent attachment of both moieties. Optical ellipsometry shows an increase of ca. 5 Å with pyrene/cholesterol co-attachment on oxidized silicon wafers, and an increase of ca. 12 Å when only pyrene was added. Steady-state fluorescence measurements indicate the presence of cholesterol reduces the efficiency of pyrene excimer formation and provides a less polar environment as sensed by the PSEDA I1/I3 band ratio. The amount of pyrene excimer formed depends on the reaction time for the adlayer co-deposition reaction. Cyclic voltammetry shows that covalently bound PSEDA is oxidized at ca. 540 mV and physisorbed PSEDA is oxidized at ca. 780 mV. AC voltammetry shows that Chol-NH2 in the adlayer reduces the electron transfer rate for the PSEDA redox reaction.  相似文献   

13.
Redox reactions with Cu(NH3)2+/1+x, Fe(CN)3?/4? and I?3/I2 were studied on |-C surfaces of MoS2 using rdes. From electrophoretic measurements, the zpc was found to occur at pH 2. The role of potential-determining ions and the effects of specific adsorption of ions on the electrode behaviour have been discussed. The change of potential with pH was ? 30 mV/pH unit for MoS2 electrode in contact with a solution saturated with molybdenum and S2? ions. From the information available, an energy level diagram has been constructed. This diagram is in agreement with the observed behaviour that electron exchange for the first three redox couples occurs with the conduction band, except that the electron transfer from the reduced species, Fe(CN)4?6 and Fe2+, in the reverse direction (oxidation) is partly and totally restricted, respectively. The charge transfer behaviour in the cse of I?3/I2 couple, appears to be much more complex, most probably involving adsorption of I? on the surface.  相似文献   

14.
Summary End-functionalized poly(isobutyl vinyl ether) (2) with a terminal amine, carboxylic acid, or ester group was prepared by quenching the HI/I2-initiated living polymer ends with ring-substituted anilines (H2N-C6H4-X, p or m; X = NH2, COOH, COOC2H5). The living polymerization of isobutyl vinyl ether and the subsequent end-capping reaction were carried out at –15°C in methylene chloride. The resulting polymers exhibited a narrow molecular weight distribution and carried one terminal function (aniline residue) per chain, according to 1H NMR structural analysis.  相似文献   

15.
The syntheses and properties of trans-[Ru(NH3)4(L)(NO)](BF4)3 (L = isonicotinic acid (inaH) (I) or ina-Tat48–60 (II)) are described. Tat48–60, a cell penetrating peptide fragment of the Tat regulatory protein of the HIV virus, was linked to the ruthenium nitrosyl through inaH. I and II release NO after reduction forming trans-[Ru(NH3)4(L)(H2O)]3 +. The IC50 values against B16-F10 melanoma cells of I and II (21 μmol L 1 and 23 μmol L 1, respectively) are close to that of the commercially available cisplatin (33 μmol L 1) and smaller than similar complexes. The cytotoxicity is assigned to the NO released from I and II.  相似文献   

16.
Poly(ethylene oxide) (PEO)/2,6-bis (N-pyrazolyl) pyridine (BNPP) polymer electrolyte based photoelectrochemical cells have been fabricated with [cis-dithiocyanato-N, N-bis (2,2′ bipyridyl-4, 4′ dicarboxylic acid)ruthenium(II)] dihydrate (N3 dye) dye complex as the sensitizer and nanoporous TiO2 film as photo anode. The introduction of 2,6-bis (N-pyrazolyl) pyridine into the poly (ethylene oxide) matrix reduces the crystallinity of the polymer and enhances the mobility of I/I3 redox couple resulting in an improved performance with a higher conversion efficiency of 8.8% in terms of light energy to electric energy when compared to that of the corresponding dye-sensitized nanocrystalline TiO2 solar cell.  相似文献   

17.
Novel uranyl acrylate complexes with general formula R[UO2(CH2CHCOO)3] (RK+, NH4+, Rb+, or Cs+) were synthesized and characterized by X-ray diffraction, IR spectroscopy and second harmonic generation (SHG) measurements. All four compounds are isostructural and crystallize in the non-centrosymmetric P213 space group. Acrylate anions act as bidentate chelating ligands forming [UO2(CH2CHCOO)3] complexes which are connected through electrostatic interactions with counter ions. SHG measurements revealed the inverted correlation between the Q = I/I SiO2 values and the radii of counter ions.  相似文献   

18.
Takuya Suzuki  Kazuhiro Hatano 《Polymer》2009,50(11):2503-3864
The gelation process and the microstructure after the gelation of polyurethane resin consisting of acryl-polyol and polyisocyanate was investigated by dynamic light scattering (DLS) as a function of stoichiometric ratio, [NCO]/[OH], and the concentration of acryl-polyol, Cpolyol. The following conclusions were obtained. First, the sol-gel transition was explained by the so-called site-bond percolation theory. Here polyol groups act as site-occupants, and isocyanate groups act as chemical cross-linkers. Second, the scattering inhomogeneities increased rather gradually around the gelation point, which are the characteristics of the gelation process not from monomeric but from oligomer units. Third, at the gelation point, the characteristic decay time of the fast mode, τfast, decreased, and the fraction of the collective diffusion mode, A, increased with [NCO]/[OH], which are due to introduction of cross-linking and/or increase of the rigidity of the network. Finally, the ensemble average scattering intensities, 〈IE, the static inhomogeneities, 〈ICE, the time-average dynamic fluctuating component, 〈IFT and the collective diffusion coefficient, D, showed remarkable dependence not only on Cpolyol but also on [NCO]/[OH]. These are due to the competition between the cross-linking and the progress of micro-phase separation.  相似文献   

19.
The presence of NO during the regeneration period of a Pt–Ba/Al2O3 Lean NOx Trap (LNT) catalyst modifies significantly the evolution of products formed from the reduction of stored nitrates, particularly nitrogen and ammonia. The use of isotope labelling techniques, feeding 14NO during the storage period and 15NO during regeneration allows us to propose three different routes for nitrogen formation based on the different masses detected during regeneration, i.e. 14N2 (m/e = 28), 14N15N (m/e = 29) and 15N2 (m/e = 30). It is proposed that the formation of nitrogen via Route 1 involves the reaction between hydrogen and 14NOx released from the storage component to form 14NH3 mainly. Then, ammonia further reacts with 14NOx located downstream to form 14N2. In Route 2, it is postulated that the incoming 15NO reacts with hydrogen to form 15NH3 in the reactor zone where the trap has been already regenerated. This isotopically labelled ammonia travels through the catalyst bed until it reaches the regeneration front where it participates in the reduction of stored nitrates (14NOx) to form 14N15N. The formation of 15N2 via Route 3 is believed to occur by the reaction between incoming 15NO and H2. The modification of the hydrogen concentration fed during regeneration affects the relative importance of H2 or 15NH3 as reductants and thus the production of 14N2 via Route 1 and 14N15N via Route 2.  相似文献   

20.
The electrochemical behavior of CoBr2 has been investigated in the presence of allylic ethers in an acetonitrile + pyridine (v/v = 9/1) mixture. Cyclic voltametry experiments show a complexation reaction between the electrogenerated cobalt(I) species and allylethers leading thus to a more stable CoI via the formation of a (η2-allyl-OR)CoIBr complex. However, this stabilization is too weak to be observed on the time-scale of cyclic voltametry in the absence of pyridine. The thermodynamic constant for the complexation has been determined. It is also shown that only one molecule of allylether is involved in the stabilisation of CoI. Interestingly, no intra-molecular oxidative addition of allylether to CoI was observed on the time-scale of a preparative electrolysis unless the reaction was carried out in an undivided cell in the presence of a sacrificial iron rod releasing iron(II) ions. Conversely, an oxidative addition has been evidenced between the electrogenerated CoI and aryl halides, ArX. The apparent rate constants for this reaction have been determined for various ArX. Importantly, the rate constants depend on the allylether concentration suggesting that the oxidative addition occurs competitively rather than consecutively with respect to complexation. More importantly, preparative-scale electrolyses of aryl halides and allylphenylether in the presence of CoBr2 as catalyst lead to the coupling product allylaryl, whereas replacement of allylphenylether with allylethylether produces only ArAr along with ArH.  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