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.
Oligomeric N-acetyl-L-glutamic acid benzyl esters with exact residue number (2, 3 or 4) have been synthesized by a stepwise procedure. The aggregational behavior of these oligomeric molecules in dioxane and benzene has been investigated by use of 1H NMR. In particular, the concentration dependence of the 1H signals for the N-terminal CH3 protons has provided evidence that an intermolecular CH3···? interaction plays a critical role in the formation of aggregates and that an intramolecular CH3···? interaction occurs in the monomolecular state.  相似文献   

2.
A series of N-acetyl-L-glutamic acid oligomeric benzyl esters with exact residue numbers (4, 6, 8, 10, and 12) has been synthesized by a stepwise procedure. It has been found that the phase maps of these oligomer-dioxane systems consist of three regions (I : an isotropic solution, II : a liquid crystalline phase, and III : a two-phase (I and II) solution). In particular, for the samples with residue numbers 6, 8, and 12, selective light scattering of coloured regions in the transparent II and III (bottom-layer) regions have been investigated by the use of UV-visible absorption spectra, leading to the conclusion that there exists a helical axis in the structure of the supramolecular aggregates formed by these oligomeric molecules.  相似文献   

3.
The method of complex-coordinate rotation is used to investigate electric-field effects on the doubly-excited 2s2p 1P0, 2p2 1De, 2p2 3Pe and 2s2p 3P0 states of He. Strong electric-field strengths up to F=0.02 Ry are used in our present study. Products of Slater orbitals are used to represent the two-electron wave functions, with lmax=8 being employed for the individual electron. Block matrices with up to Lmax=5 (H-states) are used to investigate the convergence behavior for the resonance parameters (resonance energy and width). When the external electric field is turned on, "classic" Stark effect is observed for the M=0 and M=ǃ components of the two singlet-spin states and for the M=ǃ components of the triplet-spin states. Comparisons are made with other calculations when available.  相似文献   

4.
The viscoelastic behavior of aqueous oleyldimethylamineoxide (ODAO) solutions was examined by varying the concentration of ODAO, cD, and the average degree of protonation, <!>, by adding hydrogen bromide to the system. ODAO forms long threadlike micelles in aqueous solution in spite of the absence of any additives. Therefore, the aqueous ODAO system shows pronounced viscoelastic behavior caused by entanglement between threadlike micelles as highly entangled polymer systems do. The frequency dependence of the storage and loss moduli for the system is perfectly described by that of a Maxwell model possessing only one set of relaxation time, F, and strength, GN. GN is proportional to the square of cD as observed in concentrated liner polymer solutions, while F decreases with increasing cD. On the other hand, the addition of HBr to control <!> remarkably increases F when GN is constant. This suggest that <!> alters the inner structure of the threadlike micelles; association of head groups, dimers or trimers of ODAO are formed in the micelles owing to hydrogen bonding between protonated ODAO or between protonated and unprotonated ODAO. An increase in the number density of the associated head groups in the threadlike micelles increases F. Moreover, the flexibility of the threadlike micelles in the system is not affected so much by a change in the number density of the associated head groups, since GN corresponding to the number density of entanglements in the system is kept constant.  相似文献   

5.
In the mixed micelles of an ionic surfactant (sodium dodecyl sulfate) with a nonionic surfactant (N-decanoyl-N-methylglucamide, hexaoxyethylene glycol-mono-n-decylether, and hexaoxyethylene glycol-mono-n-dodecylether), the critical mole fraction, Xic, of the ionic surfactant has been determined, below which the counterion is completely released from the micelles. The values of Xic are 0.074, 0.11, and 0.11, for the respective nonionic surfactants. The valences, i.e., the aggregation numbers of the ionic surfactant, of the mixed micelles at Xic are almost close to each other, around 6. At Xic, the critical surface charge density (about 0.03 Cm-2) for counterion condensation was tentatively calculated. In the present study, a differential conductivity method was applied.  相似文献   

6.
Copolymers of N-isopropylacrylamide (IPA) and alkyl acrylates [methyl acrylate (MA), ethyl acrylate, and butyl acrylate] or vinyl acetate have been prepared and their phase transitions in water have been observed by means of IR spectroscopy. The incorporation of these alkyl acrylates into a poly(IPA) (PIPA) chain induces a decrease in the phase-transition temperatures, Tp, and the magnitude increases with increasing size of the alkyl chains. The profiles of the C=O stretching absorption bands of the ester groups [9(C=O)ester] and the IR bands due to IPA units exhibit critical changes at the Tp of these copolymers. The 9(C=O)ester bands shift slightly toward higher wavenumbers (blueshift) upon phase transition, while the amide I and amide II bands of the IPA units undergo a blueshift and a redshift, respectively. Analysis of the 9(C=O)ester band of PIPA-MA by using a curve-fitting method shows that it consists of three components, at 1,703, 1,720, and 1,738 cm-1. The relative peak area of the largest component (1,720 cm-1) is almost constant, and those of the 1,703-cm-1 and 1,738-cm-1 components increase and decrease with increasing temperature during the phase transition, respectively. However, the changes are rather small, suggesting that changes in hydrogen bonding of the C=O groups of MA units upon phase transition are not significant. The 9(C=O)ester bands of other comonomers examined here also exhibit similar changes. The situation is consistent with the change in the hydration states of the amide groups of IPA units, most of which associate with water molecules through hydrogen bonds even after the phase separation.  相似文献   

7.
Four strong polyelectrolyte samples of 2-(acrylamido)-2-methylpropanesulfonic acid (AMPS) and N,N-dimethylacrylamide (DMAA) were radically copolymerized with a single label of naphthalene or pyrene, with both labels and without label, containing about 40 mol % AMPS. Fluorescence nonradiative energy transfer (NRET) IPy/INp, anisotropy r, I1/I3 and excimer emission IE/IM of pyrene labels were observed in dilute aqueous solutions with and without cationic surfactant of cetyltrimethylammonium bromide (CTAB). The overlap concentration was determined as 3 g/L from the appearance of intermolecular excimer. The variation of intra- and intermolecular NRET with total polyelectrolyte concentration showed that the charged chains preferentially interpenetrated each other rather than reduce their coil volume as their concentration beyond the overlap threshold. By binding with CTAB, the polyelectrolyte chain became more coiled as known from the reduced viscosity. The intramolecular NRET was dominant when [CTAB]Д᎒-5 M and then the intermolecular NRET occurred at higher CTAB concentrations with hydrophobic aggregation between CTAB tails bound on different polyelectrolyte chains. The CTAB concentration corresponding to the maxima of IPy/INp just is equal to the AMPS monomer concentration, indicating the formation of 1:1 binding between surfactant and polyelectrolyte in very dilute solutions. Added salt of NaCl up to 0.1 M hardly affected the intramolecular NRET but affected the IPy/INp value for the intermolecular NRET.  相似文献   

8.
In the micellar solution of SDS, the partition coefficient (Kx) of following branched alkanols at infinite dilution was determined by applying a differential conductivity method: the alkanols used were i-CmH2m+1OH (m=4-9, i=1-5) in which the position of OH group (i) shifts from an end to the center of a hydrocarbon chain. The method provides two significant quantities, d!/dXam and dCsf/dCaf in addition to Kx. The following results have been obtained. (1) The dependence of Kx on i indicates that the hydrophobicity of alkanol is weakened with increasing i, whereas the increase in m strengthens the hydrophobicity. (2) The degree of counterion disossiation of micelles (!) is accelerated by the solubilized alkanols in micelles (mole fraction: Xam) and the acceleration rate, d!/dXam (=0.17), depends on neither m nor i. (3) In the bulk water, the monomerically dissolved alkanols (concentration: Caf) depresses the concentration of free monomer surfactant (Csf), and the depressing rate, dCsf/dCaf, in micellar solution is identical with the corresponding quantity, ((CMC/(Ca)o at CMC.  相似文献   

9.
The effect of a water-soluble uncharged polymer on the stability of the lamellar phase of the Aerosol OT (AOT)/water system is studied. The lamellar phase still exists when water is replaced by an aqueous solution of poly(N,N- dimethylacrylamide) (RgƼ᎒2 Å). Since the coil dimensions are (much) larger than the thickness of the water layers (dwᅣ Å), the polymer molecules do not enter the lamellar phase. Instead segregation in small domains occurs, and in equilibrium with the AOT-rich phase another separate phase containing the polymer is formed. The polymer-rich phase exerts an osmotic pressure that reduces the water content in the AOT-rich phase, and by compression the repeat distance is reduced.  相似文献   

10.
The influence of charged components of cationic polyelectrolytes on the dewatering of clay-containing suspensions was investigated with a view to better predicting the efficiency of flocculating agents. In flocculation and dewatering experiments on suspensions of harbour sediment and gravel washings, flocculating agents of the polyacrylamide-co-(trimethylammoniumpropyl chloride) (PTCA) type exhibited different dewatering efficiencies depending on the degree of cationicity, F. For harbour sediment, a dewatering index, ID, of 80 was achieved with the highly charged PTCA 3 (F=40%) at 30% lower flocculant dose than with the weakly cationic PTCA 1 (F=3%). However, for gravel washings PTCA 1 proved to be more effective: for comparable degrees of dewatering (ID=80) approximately 40% less flocculant was required than for PTCA 3. In shear experiments on gravel washings and model suspensions with particles of differing size (d50=0.5 und 5.7 µm) weakly cationic PTCA 1 exhibited an increased floc stability at lower concentrations than is necessary to achieve maximum ID values. For harbour sediments and model suspensions with unimodal particle size distributions this stability did not occur until the doses used were higher than the concentrations needed to achieve maximum ID values.  相似文献   

11.
A wide investigation of the solubilization of the water-soluble salt Co(NO3)2 in water/sodium bis(2-ethylhexyl)sulfosuccinate (AOT)/n-heptane microemulsions and of some physicochemical properties of the Co(NO3)2/AOT/n-heptane and Co(NO3)2/AOT systems has been carried out. After saturation of water/AOT/n-heptane microemulsions with pure Co(NO3)2, the Co(NO3)2/AOT composites were prepared by complete evaporation under vacuum of the volatile components (water and apolar solvent) of the salt-containing microemulsions. It was observed that these composites could be totally dissolved in pure n-heptane, allowing the solubilization of a noticeable amount of Co(NO3)2 in a dry apolar organic solvent. By UV-vis-near-IR spectrophotometry, some information on the state of Co(NO3)2 within water-containing or dry AOT reversed micelles was acquired, whereas by small-angle X-ray scattering it the occurrence of small nanoparticles in the salt-containing dry AOT reversed micelles was ascertained. Surprisingly, the analysis of the X-ray diffraction spectra corroborated by UV-vis and X-ray photoelectron spectroscopy data of Co(NO3)2/AOT composites led to the hypothesis that nanoparticles are mainly constituted of sodium nitrate resulting from the metathesis reaction between AOT and Co(NO3)2. By transmission electron microscopy, information on the size and the size distribution of the nanoparticles in salt/AOT composites was gained.  相似文献   

12.
The values of the critical micelle concentration (cmc) and the degree of electrolytic micelle dissociation, a, for sodium dodecyl sulphate (SDS) as a function of the concentrations of the electrolytes added, NaCl, KCl, NaF, NaClO4, NH4ClO4, and Mg(ClO4)2, have been determined. The values of the SDS cmc have been shown to depend on the kind and concentration of the electrolyte cations. The electrolyte cations cause a decrease of the cmc in the following order: Na+4++2+. Moreover, a depends on the kind and concentration of the electrolyte added. The electrolyte anions have a much smaller effect on the values of a than the cations. The anions enhance a in the following order: F->ClO4->Cl-. The effect of different electrolyte cations on a is observed; moreover the values of a either increase or decrease with the electrolyte concentration. Other micellization parameters of SDS versus the concentration of the electrolytes added have been calculated.  相似文献   

13.
Amphiphilic polymers consisting of a statistical distribution of octadecyl methacrylate (ODMA) and acrylic acid in respective molar ratios of 83-22 and 17-78 mol% and in a molecular-weight range of 2.35-4.70᎒4 gmol-1 have been synthesized. The series of polymers consisting of various mole fractions of ODMA and acrylic acid are expected to exhibit unique characteristics resembling ionomer to hydrophobically modified polyelectrolytes. The changes in the I3/I1 emission intensity ratios of pyrene, occurring in the presence of tetrahydrofuran (THF) solutions of the polymers have been taken as the main basis for inferring solution structures. The polymers are found to form random-coil to collapsed-coil/aggregated structures in THF solvent depending on the copolymer compositions. The polymer consisting of 83 mol% ODMA and 17 mol% acrylic acid behaves as an ionomer, capable of forming collapsed-coil structures at concentrations of 0.02 gml-1 and above as shown by a very high I3/I1 of 1.20 (I3/I1 of pyrene in THF is 0.85). In contrast, the poly(octadecyl methacrylate) homopolymer and the sets of copolymers consisting of a very high proportion of acrylic acid to an extent of 73 mol% and above contribute to almost negligible or very small changes in I3/I1 similar to the homopolymer, poly(octadecyl methacrylate), suggesting the formation of random-coil structures.  相似文献   

14.
A series of N-acetyl-L-glutamic acid oligopeptide benzyl esters with exact residue numbers 4, 6, and 12 has been synthesized by a stepwise procedure. For these oligopeptide–dioxane binary systems, the behavior of the liquid-crystalline phases has been examined␣by the use of 2H NMR, and the results indicate that highly ordered aggregates formed by these oligopeptides in dioxane are in an alignment similar to that in a nematic mesophase. Received: 27 February 2001 Accepted: 8 May 2001  相似文献   

15.
Films of copper sulfides of varying composition are formed in a surface matrix of polyamide by a sorption-diffusion method using solutions of higher polythionic acids, H2SnO6 (n>6), as sulfuring agents. A film of nonstoichiometric CuxS (x=1.06-1.95) is obtained when the sulfured polyamide is treated with a solution of Cu(I-II) salt. The value of x in CuxS decreases with the prolongation of the period of polyamide sulfuration in the H2SnO6 solution and increases with the prolongation of the period of sulfured polyamide interaction with the copper salt solution. The films obtained are formed from two main phases: yarrowite (Cu1.12S) and anilite (Cu1.75S). Depending on the polyamide sulfuration and the sulfured polyamide interaction with a solution of Cu(I-II) salt conditions, CuxS films on polyamide of different electrical conductivity were obtained. The sulfide with a composition close to CuS has the highest electrical conductivity.  相似文献   

16.
The mixed micelle formation by hexadecyltrimethylammonium bromide (HTAB), hexadecylpyridinium bromide (HPyBr), and hexadecylpyridinium chloride (HPyCl) with benzylhexadecyldimethylammonium chloride (BHDACl) was studied with the help of conductivity, 3, and viscometric measurements. Each 3 curve for HTAB, HPyBr, and HPyCl showed a single break, whereas double breaks were observed in the case of BHDACl and its binary mixtures with HTAB, HPyBr, and HPyCl. The first break, c1,M, was due to mixed-micelle formation, whereas the second one, c2,M, was due to the structural micellar transitions in the mixed micelles formed during the first break. The quantitative analysis of the mixed-micelle formation corresponding to the first break was carried out with the help of regular solution and Motomura's approximations. The BHDACl plus HTAB and BHDACl plus HPyBr mixtures showed weak synergistic interactions, whereas the BHDACl plus HPyCl mixture was close to ideal. A variation in the degree of dissociation corresponding to the first and second breaks demonstrated a similar variation over the whole mixing range and this was attributed to the identical nature of synergism in both kinds of micellar aggregates available at c1,M, and c2,M. They were rich in BHDACl over most of the mole fraction range. Unlike conductivity measurements, the relative viscosity demonstrated that the micellization of HTAB, HPyBr, and HPyCl proceeded through a minimum, whereas that for BHDACl and its mixtures showed a strong maximum in each case, being stronger in the case of the BHDACl plus HTAB and BHDACl plus HPyCl mixtures.  相似文献   

17.
Langmuir films of naphthenic acids at different pH and electrolyte concentrations are reported. The polydisperse naphthenic acids were commercially available, while two single-component naphthenic acids [5#(H)-cholanoic acid and 1-naphthalenepentanoic acid, decahydro- (9CI)] were synthesized. 1-naphthalenepentanoic acid, decahydro- is too water-soluble to form stable monolayers. 5#(H)-Cholanoic acid and Fluka naphthenic acid form stable films when cations are present in the aqueous subphase. At lower pH the cations are less influential since the naphthenic acids are not protolysed and metal naphthenates cannot be formed. pKas for 5#(H)-cholanoic acid is determined to 5.65. The micellisation of the naphthenates at high pH is described.  相似文献   

18.
Photo fragmentation studies of stored mass selected metal cluster ions of a large size range are reported. The experimental method and the data evaluation are described in detail. Gold cluster ions were produced by laser vaporization and stored in a Penning trap. After size selection they were electronically excited by irradiation with a pulsed laser beam. Relaxation by evaporation of neutral atoms and dimers was observed as a function of photon energy. From these data upper and lower limits for dissociation energies are determined for Au+n (n=3 to 23).  相似文献   

19.
The photocontrol of polyion complex formation between polylysine bromide (PlysBr, Mw=21,800) and sodium chondroitin sulfate (NaChs, Mw=24,700) in aqueous solution was examined in the presence of a photochromic dye, pararosaniline leucohydroxide (leuco(OH)), with special reference to its pH-dependence. Under pH conditions examined (6< pH <12), NaChs was almost fully dissociated, while dissociation of PlysBr reduced with increasing pH, resulting in a random coil-to-!-helix transition at around pH 9.4. Thus, the complex formation clearly showed pH-dependence. On the other hand, leuco(OH) undergoes photodissociation to increase the solution pH. By coupling these processes, we aimed to control the complex formation by photoirradiation. Turbidimetry, dynamic light scattering and electrophoretic light scattering measurements showed that the optimum condition for the complex formation was at approximately pH 10.5, where the net charge of complex was nearly zero, for solutions of [NaChs]=1.0 mM, [PlysBr]=1.5 mM (in molalities of dissociable groups), at a salt concentration of 1.5 mM. In the presence of 0.5 mM leuco(OH), the variation of complex size was clearly observed by photoirradiation. The hydrodynamic radius of the complex stayed almost constant at pH <9 and pH> 11, increased at 9< pH <10, and decreased at 10< pH <11. The observed trend could be reasonably explained in terms of the pH change due to photodissociation of leuco(OH). A photoinduced coil-to-!-helix transition of PlysBr in the presence of leuco(OH) was also reported.  相似文献   

20.
Well-defined hydroxy end-functionalized poly(n-butyl acrylate)s (PBA-OH and PBA-(OH)2), were prepared by atom transfer radical polymerization (ATRP) and used as reactive stabilizers for the preparation of polyurethane in dispersed medium. PBA-OH was obtained by end-capping the growing poly(n-butyl acrylate) chains with allyl alcohol added in excess at the end of the polymerization. The two hydroxyl functions of PBA-(OH)2 were fixed at one end of the poly(n-butyl acrylate) chains either by initiation or by chain-end functionalization reactions. The latter were protected under the form of cyclic acetal and attached either to the initiator bearing a secondary bromine or to the terminating agent carrying a poorly reactive vinylic unsaturation. PBA-OH and PBA-(OH)2 have been successfully used as reactive stabilizers (surfmers) to prepare core-shell polyurethane particles in dispersed medium. The final particle size was found to be very much dependent to parameters such as the molar mass, concentration and valence of the reactive stabilizer as well as the manner of addition of the reactants during the procedure.  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