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.
Rod-like molecules with a strong dipole in the long axis were mixed with synthetic polyisoprene (IR305) in low concentrations (0.1–5 wt%) and measured dielectrically in the frequency range 10–2–106 Hz and the temperature range –60°–0°C (glass relaxation range). The measurements showed an additional symmetrical relaxation process slower than the glass process by a factor ranging from 10 to 100 times, depending on the length and structure of the label. The label relaxation process was characterized by a narrow distribution of relaxation times and its strong coupling with the segmental motion. For the labels with only a longitudinal dipole moment, no relevant part of the label relaxation was observed in the relaxation frequencies of the glass process; this may be attributed to the parallelism and compactness of the polyisoprene chains. In contrast, the labels with an additional vertical dipole component showed an additional relaxation process superimposed on the glass relaxation with also almost the same distribution of relaxation times. The slow label relaxation could be attributed to a diffusion or rotational diffusion relaxation process through a multistep relaxation mechanism determined by the segmental motion of the matrix.  相似文献   

2.
Polystyrenes with different concentrations of side groups with cyano groups were prepared and complex dielectric constants were measured in the range of the glass transition temperature and the frequency range of 10–2–107 Hz.The GPC and DSC measurements showed that the molecular weight of these polystyrenes was about 10500 g/mole and the glass transition temperatures were 89.5°C for all samples.The dielectric relaxation spectra obtained for the side group polystyrene labels and also the chain-end polystyrene labels prepared before [9] were analyzed to find out the degree of coupling of the chain-end and side-group labels with the cooperative reorientation of the polymeric matrix. The analysis of the spectra was carried out using the analysis method developed by Mansour and Stoll [6].The results obtained showed that both end- and side-group labels are strongly coupled with the segmental reorientation and relax with relaxation times longer than that of the segments.The value of logf m = (logf m(label)) – logf m(matrix)) was obtained from the recently designed comparison diagram suggested by Mansour and Stoll [6, 14]. The value of logf m depends on the label length in the case of chain-end labels.It was surprising to find that the side groups relax slower than the segments by only 0.9 decades. These results obtained implied that the label relaxes through a multistep relaxation mechanism of the side and end groups and not through a diffusion mechanism of the whole chain. In addition, the effective lengths of the relaxing units were determined using the empirical equation obtained before in the case of rodlike molecules in polyisoprene [7].  相似文献   

3.
A study has been undertaken of stress relaxation in ovalbumin thermotropic gels with a concentration of 8–20%, depending on time and temperature of heating (respectively, 20–60 min, 70°–110°C), at pH 2.5–10.0. In all instances, the dependence of the initial gel elasticity modulus on heating has a single maximum. Gelation conditions corresponding to this maximum are considered optimal. Optimal gelation time is 30 min, regardless of pH. On the other hand, the optimal heating temperature depends on pH. To the right and left of the isoelectric point of protein (2.5pH<4.0 and 5.5G) of gels on heating conditions, pH and protein concentration (X 1,X 2,X 3,X 4), as well as on time of relaxation (t) may be generally described asG(X 1,X 1,X 1,X 1,t)=G e(X 1,X 2,X 3,X 4)f(t), whereG e is the equilibrium value of the elasticity modulus, and f(t) the relaxation function. Thus, a change in the parameters only affects the value of the equilibrium elasticity modulus, and exerts no effect on the relaxation time spectrum. For this reason, all the relaxation curves obtained may be transformed into two normalized relaxation functions:f(t)=f(t)/f(1)=G(X 1,X 2,X 3,X 4,t)/G(X 1,X 2,X 3,X 4, 1)Each of these normalized functions corresponds to one of the homologous groups. Rheological similarity of gels in each homologous group evidently points to their structural similarity. Invariance of the gel relaxationproperties with regard to protein concentration, leads to a concentration dependence of the equilibrium modulus at various pH values. These dependences are curvilinear on a double logarithmic scale. The slope of the curve exceeds 2 in the entire concentration interval studied. In other words, the dependences obtained cannot be described by the usual law of squares. On the other hand, they adequately match Hermans theoretical relation for a network formed by random association of identical polyfunctional particles without cyclization. This simple model evidently gives a true picture of the major regularities of thermotropic gelation for ovalbumin. An agreement between this theory and experiment was achieved for a protein concentration ofC *=6.0±1.0% at the gel point regardless of pH. Invariance of gelpoint position with regards to pH demands further confirmation.List of symbols T h,t h heating temperature and time - T h * ,t h * optimal heating temperature and time - C protein concentration - C * protein concentration in gel-point - G relaxation modulus - G e equilibrium modulus - f(t) relaxation function - t time of relaxation - f(t) normalized relaxation function - fT A (t), f B (t) normalized relaxation functions of groups A and B - G 1 T h,t h-reduced modulus - G 2 T h,t h, pH-reduced modulus - G 3 C-reduced modulus - b 1 T h, th reduction parameter of modulus - b 2 pH reduction parameter of modulus - b 3 C reduction parameter of modulus - Wg gel-fraction  相似文献   

4.
Grafting reactions of phenyl groups on silica substrates using as reagents various phenylsilanes with variable distances between the silicon atoms and the aromatic ring by 0, 1, and 2 methylene groups [Si–(CH2) n –C6H5] were studied. Two different silicates have been selected as source of silica: sepiolite and tetraethyl-orthosilicate (TEOS). Sulfonic- and nitro-derivatives prepared from these phenyl compounds by electrophilic substitution reactions, have been obtained. The mechanism of these processes has been studied in relation to the number of methylene separating groups belonging to the starting reagents. The characterization of such systems has been achieved by XRD,29Si and13C NMR/MAS, IR, and laser microprobe mass spectrometry.Parts I, II and III published in this Journal (1978) 256:135–139, (1979) 257:178–181 and (1985) 263:1025–1030  相似文献   

5.
An isotactic polypropylene film was stretched at 120 °C in poly(ethylene glycol) and thermally shrunk at various temperatures. Proton spin-lattice,T 1, and spin-spin,T 2, relaxation times were measured using a broad line pulse spectrometer operating at 19.8 MHz in the temperature range 40 °C–100 °C. The temperature ofT 1 minimum shifts to higher temperatures and the value ofT 1 minimum increases in magnitude as the stretching ratio is increased. In contrast the temperature ofT 1 minimum shifts to lower temperatures as shrinkage is increased, whereas the value ofT 1 minimum increases in magnitude because of the increase in crystallimty during shrinkage. T2a, the longestT 1 associated with the mobile amorphous regions, increases during shrinkage, indicating that chain mobility in the amorphous regions increases substantially during shrinkage. It was found that an orientation function of the amorphous regions,f a, correlates well withT 2a .Presented in part at the 52nd Annual Meeting of the Japan Chemical Society, Kyoto, April 1986.  相似文献   

6.
Mechanical relaxation processes in polymer melts and networks are discussed. This is performed by decomposing master curves of the dynamic shear compliance into i) glass relaxation with its plateau complianceJ eN ; ii) shearband process with its relaxation strengthJ B , which is reciprocal to the total crosslink densityp c ; and iii) flow relaxationJ F and viscous flow (for uncrosslinked melts only). Plateau complianceJ eN > is exponentially reduced only by effective crosslinks (p c * p c /30). This behavior is understood on the level of a meander superstructure, which includes shearbands. The observed saturation inJ eN at higher dicumylperoxide (DCUP) crosslinking-which doesn't appear with radiation-can be explained by the lack of chemically induced effective crosslinks across the interfaces among meander cubes. This lack may be a consequence of DCUP molecules concentrating at the interfaces and thereby preventing the contact and radical recombination between chains at adjacent meander faces.Crosslink densitiesp c (per monomer), determined from the reduction of shearband relaxation strength, vary linearly with the crosslinking agent and read: pc2.4 · 10–2 Dose/MGy andp c 0.97 · 10–2 DCUP/phr for radiation and DCUP crosslinking, respectively. This implies, e.g., that a dose of 0.4 MGy (40 Mrad) is equivalent to 1 part DCUP phr in a crosslinking polyisoprene. From activation-curve analysis it follows that3 r/d stays constant, and s - so (free energy of formation of a segment-dislocation) andQ y -Q yo (activation energy for segmental jumps) vary with the square ofP c , as does the glass temperaturT g -T go from DSC measurements.  相似文献   

7.
Electrical and dielectrical properties of poly(bis(-phenoxyethoxy)phosphazene) (I) and its complexes with various content ratios of AgSO3CF3 to monomeric unit (0.25/1 (II) and 0.5/1 (III) in molar ratio) were investigated.Dc conductivity of respective samples at 18 °C were 6.1×10–12, 4.4×10–9, and 7.1×10–8 S/m.Dc conduction was considered to be due to ion hopping. Charge mobility ranged from 3×10–12 to 6× 10–11 m2/Vs depending on the applied field in sample II. In sample I, a tan peak was found which can be ascribed to molecular relaxation of main chains. The peak vanished upon introducing AgSO3 CF3. Temperature dependence of total conductivity ( T ) measured byac method in the temperature range between –150 °C and 50 °C showed several peaks at the temperatures corresponding to the peak temperatures of tan. Total conductivities of respective samples at 100 kHz were 4.9×10–7 (69 °C), 1.7×10–4 (45 °C), and 1.5×10–4(40°C)S/m.  相似文献   

8.
Ball-like molecules with strong dipoles (labels) were mixed with synthetic polyisoprene (IR305) in low concentrations (<1%) and measured dielectrically in the frequency range 10–2–107 Hz and the temperature range –70–0°C (glass relaxation region). Calorimetric measurements showed that this type of label has a plasticizing effect on the polymeric matrix. The dielectric measurements showed that these ball-like molecules relax through cooperative rotations with the polymeric segments and at the same relaxation frequency. In addition, the label molecules showed a high-frequency local relaxation process. The relaxation strength ratio of the local process (X local) to the total relaxation strength of the label was found to be dependent on the volume as well as on the shape of the label. A comparison between the relaxation behaviors of the ball-and rod-like molecules, having the same volume, showed that the length of the label is also an important parameter for the determination of the local contribution as well as of the cooperative relaxation mechanism of the label. The label relaxation process is discussed in relation to the molecular packing of the host polymer.  相似文献   

9.
To continue dynamicc p measurements in the range of smallest temperature rates and non-linear thermal relaxation investigations into the linear range, simultaneousc p and thermal relaxation measurements were carried out in an adiabatic vacuum calorimeter, using the pulse heating method. The rate-dependentc p behaviour, known from dynamic measurements, does not continue at small temperature rates. This is confirmed by the relaxation process which is observed. The results suggest an extended interpretation of the glass transition in atactic polymethylmethacrylate (PMMA).  相似文献   

10.
Measurements of the complex permittivity of crosslinked polyurethanes at different temperatures in the frequency range 1–105 Hz are discussed with respect to shape of relaxation curves. Using a new model (published in preceding paper) the shape parameters are related to small and large scale interaction.  相似文献   

11.
Monolayer properties of irisresorcinol [5-(cis-10-heptadecenyl) resorcinol] were measured at the air/water interface. TheA-T isobars of the monolayers at 10 and 15 mN/m gave two-dimensional thermal expansivities of 1.4 × 10–4/K and 1.3 × 10–4/K at a temperature span from 7–40 C, respectively. The- A isotherms of the material showed only a little dependence on temperature from 5–35 C and onpH except at highpH, where monolayers expanded by ionization of resorcinol headgroups. Some types of saccharose in the subphase exhibited a characteristic interaction with irisresorcinol in monolayers, and there is a possibility that this material will be used for molecular recognition of some saccharoses.  相似文献   

12.
The structure-property relationships derived here permit the prediction of both the zero-shear viscosity 0, as well as the shear rate dependent viscosity . Using this molecular modeling it is now possible to predict over the whole concentration range, independently of the molecular weight, polymer concentration and imposed shear rate. However, the widely accepted concept: dilute — concentrated, is insufficient. Moreover it is necessary to take five distinct states of solution into account if the viscous behavior of polymeric liquids is to be described satisfactorily. For non-homogeneous, semi-dilute (moderately concentrated) solutions the slope in the linear region of the flow curve (= must be standardized against the overlap parameterc · []. As with the 0-M-c-relationship, a-M -c- relationship can now be formulated for the complete range of concentration and molecular weight. Furthermore, it is possible to predict the onset of shear induced degradation of polymeric liquids subjected to a laminar velocity field on the basis of molecular modeling. These theoretically obtained results lead to the previously made ad hoc conclusion (Kulicke, Porter [32]) that, experimentally, it is not possible to detect the second Newtonian region.Roman and Italian symbols a exponent of the Mark-Houwink relationship - b exponent of the third term of the 0-M -c relationship - c concentration /g · cm–3 - E number of entanglements per molecule - F(r) connector tension - f function - G i A shear modulus; A indicates that it /Pa has been evaluated by a transient shear flow experiment; i is the shear rate to whichG A refers to - G storage modulus /Pa - G p plateau modulus /Pa - H() relaxation spectrum /Pa - h shift factor (0/r) - K H Huggins constant - K b third constant of the 0-M -c relationship - K constant of the Mark-Houwink relationship - M molecular weight /g · mol–1 - M e molecular weight between two /g · mol–1 entanglement couplings - N number of segments per molecule - n slope in the power-law region of the flow curve - p p-th mode of the relaxation time spectrum - R gas constant /8.314 J·K–1·mol–1 - r direction vector - T temperature /K Greek symbols ß reduced shear rate - shear rate /s–1 - shear viscosity /Pa·s - s solvent viscosity /Pa·s - sp specific viscosity - 0 zero-shear viscosity /Pa·s - apparent viscosity at shear rate - reduced viscosity - viscosity of polymeric liquid in /Pa·s the second Newtonian region - [] intrinsic viscosity/cm3·g–1 - screening length/m - /g·cm –3 density - relaxation time/s - 0 experimentally derived relaxation time/s - angular frequency of oscillation Indices conc concentrated - corr slope corrected - cr critical - deg degradation - e entanglement - exp experimental - mod moderately concentrated/semi-dilute - n number average - p polymer - R Rouse - rep reptation - s solvent - sp specific - theo theoretical - weight average - relaxation time - o experimental or steady state - * critical - ** transition moderately conc. — conc. - + transition dilute — moderately cone. Paper presented at the 2nd bilateral U.S.-West German Polymer Symposium, Yountville, the 7th–11th September 1987.  相似文献   

13.
Steric repulsion of polyoxyethylene groups for emulsion stability   总被引:1,自引:0,他引:1  
Rapid coalescence was studied on liquid paraffin emulsion stabilized with a series of poly(oxyethylene) dodecyl ethers [C12H25 (EO),n=1, 2, 3, 4, 5, 6, 7, 8] and of poly(oxyethylene) nonylphenyl ethers [C9H19(EO) n ,n=2, 4, 5, 6, 12]. The turbidity of emulsion was measured as a function of the solution pHs at constant ionic strength of 0.1 mol dm–3.As a result, it was found that the emulsions (which were formed with C12H25(EO) n surfactants having less than four oxyethylene groups, or with C9H19 (EO) n surfactants having less than six oxyethylene groups) brought about rapid coalescence in the bulk pH between 2.03.5, which corresponded to the zero point of charges for the emulsions of the present systems. According to the Tadros treatment for emulsion flocculation, the total flocculation potennual was estimated as a function of the distance relative to the number of oxyethelene groups in the surfactants. The critical coalescence energy was obtained as –343 ×10–19 J for the C12H25(EO) n surfactants and –2.14×10–19) J for the C9H19 (EO) n surfactants. Furthermore, the formation of a hole for coalescence was estimated under the simple assumption that the coalescence was caused only by the energy dissipation.  相似文献   

14.
The study was extended to analysis of mass, size and conformation of micelles formed in aqueous solutions of ethoxylated nonyl phenols. The results obtained by ultracentrifugal technique between 293 and 323 K have proved that the slightly ethoxylated nonyl phenols form micelles with high molecular mass and larger size at constant temperature, while the increasing length of the ethylene oxide chain favours formation of micelles of smaller molecular mass and size. The transformation of conformation from oblate to spherical shapes ensues with increasing temperature at constant ethoxy number or with ethoxylation at constant temperature. The second virial coefficient decreases with increasing temperature and decreasing ethoxy number. In accordance with the earlier conclucions, the change of the second virial coefficient relates to enhanced variation of monomer solubility, stabilization of micelle structure and increased deviation from ideal behaviour of a given micellar system.Symbols a major axis of micelle, Å - a m attractivity factor, cm3 erg molecule2 - b minor axis of micelle, Å - c concentration, g dm–3 - c b equilibrium concentration at the bottom of the cell, g dm–3 - c m equilibrium concentration at the meniscus of the cell, g dm–3 - c o initial concentration in the cell, g dm–3 - c M critical micellization concentration, mol dm–3 - e eccentricity - f IS Isihara-constant - f/f o frictional ratio of micelle - amount of water in micelle per ethoxy group, mol H2O/mol EO - n aggregation number, monomer micelle–1 - n EO number of ethoxy groups - r distance of Schlieren peak from the axis, cm - r b distance of cell bottom from the axis, cm - r m distance of cell meniscus from the axis, cm - R h equivalent hydrodynamic radius of micelle, Å - s t sedimentation coefficient, s - reduced sedimentation coefficient, s - reduced limiting sedimentation coefficient, s - ¯v t volume of micelle, cm3 micelle–1 - partial specific volume of solute, cm3g–1 - partial specific volume of solute reduced to 293 K, cm3 g–1 - B a, Be constants, cm3 mol g–2 - B 2 second virial coefficient, cm3 mol g–2 - M m a mass average apparent molecular mass of micelle, g mol–1 - M m mass average molecular mass of micelle corrected withB 2, g mol–1 - M m cM mass average molecular mass of micelle belonging toc M, g mol–1 - M 1 mass average molecular mass of monomer, gmol–1 - N A the Avogadro's number, molecule mol–1 - R universal gas constant, erg mol–1 K–1 - T temperature, K - t o dynamic viscosity of solvent atT temperature, g cm–1 s–1 - dynamic viscosity of solvent at 293 K, g cm–1 s–1 - t density of solution atT temperature, g cm–3 - t o density of solvent atT temperature, g cm–3 - density of solvent at 293 K, g cm–3 - angular velocity, rad s–1 - time, s  相似文献   

15.
Ball-like molecules with strong dipoles (labels) were mixed with technical polystyrene (PS168N) in low concentrations (<0.5% wt) and measured dielectrically in the frequency range 10–2–107 Hz, and the temperature range 100°–135°C (glass relaxation region). The measurements showed that these ball-like molecules relax cooperatively with the polymeric segments with relaxation times lying at the high-frequency tail of the glass process. The activation energy of the main label process is found to be very similar to that of the glass process of the polystyrene segments and also has the same temperature dependence. This finding implies the existence of an additional mode of relaxation in the dielectric spectrum of the glass process of polystyrene (compared to polyisoprene). Considering the different behavior of the ball-like molecules in polystyrene and polyisoprene and the temperature dependence of the half-width of dielectric loss peak in different polymers, we suggest that the polymers could be classified into three classes according to the available dielectric relaxation modes in the glass process. In addition, the label molecules showed a high-frequency local relaxation process. The relaxation strength ratio of the local process (X local) to the total relaxation strength of the label was found to be dependent on the volume of the label. This phenomenon could supply a new method for the determination of the mean size of the holes (voids) representing the free volume of the host matrix.  相似文献   

16.
Contact angles, measured with various liquids, have been employed to calculate the surface free energies of glass after adsorption of quaternary ammonium chlorides with a variable hydrocarbon chain length 8n16. The thickness of the adsorbed layers has been determined ellipsometrically. A clear relation is observed between the measured parameters and the hydrocarbon chain lengthn, if only the extremesn=8 andn=16 are considered. Surface free energies decrease from 138 erg.cm–2 for clean glass to 101 and 64 erg.cm–2 forn=8 andn=16, respectively, at the highest concentration tested (7.5 mM). The adsorbed layer thickness of C8 amounts to approximately 50 % of the thickness observed for C16. No clear relation between the measured parameters is observed for the intermediate hydrocarbon chain lengths, which presumably reflects the many configurations possible in these adsorbed layers. It is envisaged that adsorption of C8 as well as C16 is restricted to a monolayer, which is completed at approximately 2 mM. In the case of C8 electrostatic repulsion between the polar headgroups will inhibit further adsorption, whereas in the case of C16 the van der Waals attraction from the adsorbed layer and the glass will probably not be sufficient to stimulate further adsorption.  相似文献   

17.
Barium ethyl(alkyl)phosphates, as new simple surfactants ((C2H5O)(RO)-PO 2 )2Ba2+ with various chain length ofR, were synthesized. The infrared spectra in the CH stretching region were measured for these surfactants in the solid state and in aqueous solution, and assignments were made. In particular, the ordering and environment of octyl chains in the different phases of the barium ethyl(octyl)phosphate-water system were studied by the Fourier-transform-infrared and Raman spectra. The CH stretching bands in the infrared spectra reflected the ordering and environment of octyl chains in each phase. The Raman band connected to the PO 2 symmetric stretching mode was sensitively shifted. This was caused by the change of aggregation structures with different Ba2+...PO 2 interaction. The infrared band arising from the PO 2 antisymmetric stretching mode was insensitive to the phase structures. The C–C stretching region in the infrared spectra was used to discuss the ordering of each phase.  相似文献   

18.
The viscoelasticity has been measured for aqueous solutions of tetradecyl-and hexadecyltrimethylammonium salicylates (C14TASal, C16TASal). The aqueous solutions of C14TASal without salt displayed the gel-like behavior at 10.0×10–2 g cm–3, but those more dilute than 3.2×10–2 g cm–3 presented the viscoelasticity similar to that of a Maxwell liquid. The Maxwell-like behavior was converted to the polymer-like one on the addition of (0.1–0.2) M NaBr or (0.02–0.2) M NaSal. The gel-like viscoelasticity can be connected with the spinnability of cohesive fracture failure, and the Maxwell-like and polymer-like viscoelasticities are concerned with the spinnability of ductile failure. The gel-like and Maxwell-like viscoelasticities originate in the pseudo-network formed by the pseudo-linkages between rodlike micelles, while the polymer-like viscoelasticity is caused by the entanglement of long rodlike micelles in semidilute and concentrated solutions. The aqueous solutions of C16TASal behaved very similar to those of C14TASal.  相似文献   

19.
Investigations on free radical copolymerization of 1-vinyl naphthalene (1-VNph, monomerM 2) with styrene (St), methyl methacrylate (MMA) and acrylonitrile (AN) (monomersm 1) in bulk at 60°C with AIBN as initiator are presented. Relative reactivity ratios were calculated by the Kelen-Tüdös method yielding:r st=0.70 ±0.23 andr 1–VNph=2.02 ±0.40 for system St/1-VNph;r MMA=0.32 ±0.10 andr 1–VNph=0.57 ±0.07 for system MMA/1-VNph andr AN=0.11 ±0.03 andr 1–VNph=0.45 ±0.09 for system AN/1-VNph.Q, e values for 1-VNph according to Alfrey, Price scheme were calculated toQ 1–VNph=1.02,e 1VNph=–0.62.  相似文献   

20.
Zwitterionic amphiphiles of the general formula H(CH2)y + N(CH3)2(CH2)n PO2C6H 5 , where the number of intercharge methylenesn is varied, were studied in dilute aqueous solution. Their critical micellar concentrations show a peculiar variation withn, first increasing asn varies from 1 to 4 and then slowly decreasing as methylenes are added up to 10. This behavior is interpreted as being the consequence of two opposite contributions. The first is the classical CMC lowering due to the increase of hydrophobic character with the total number of methylene groups in the surfactant molecule. The second contribution is the increase in the dipole moment of the zwitterionic headgroup withn, leading to stronger dipole-dipole repulsions between headgroups at the micellar surface. Experimental results suggest that the dipole moment does not increase linearly withn because of the polymethylene chain flexibility. This is supported by13C NMR relaxation experiments.  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