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.
Laser light scattering including angular dependence of total integrated scattered intensity and of the spectral distribution has been used to characterize five samples of poly(1,4-phenylene terephthalamide), PPTA (commercially known as Kevlar), of different molecular weights in 96% sulphuric acid and 0.1 NK2SO4. The data are supplemented by intrinsic viscosity measurements used to detect the possible effects of association, by differential refractometry providing a measure of the refractive index increments in mixed solvents (H2O, H2SO4 and K2SO4) and by spectrophotometry for the extinction coefficient needed in the correction of attenuation in light scattering studies. The results show 〈DZ = 2.11 × 10?5M?W?0.75cm?2s?1 in reasonable agreement with an average of many of the published intrinsic viscosity data obeying [η] = 1.09 × 10?3 Mw1.25 ml g?1 and w expressed in g mol?1.  相似文献   

2.
By using laser light scattering (LS) and size exclusion chromatography combined LS, we have investigated the molecular weight and chain conformation of amylopectin from rice of India (II‐b), japonica (IJ‐b), and glutinous (IG‐b) in dimethyl sulfoxide (DMSO) solution. The weight‐average molecular weight (Mw) and radius of gyration (〈S2½) of amylopectin were determined to be 4.06 × 107 and 128.5 nm for India rice, 7.41 × 107 and 169.6 nm for japonica rice, 2.72 × 108 and 252.3 nm for glutinous rice, respectively. The 〈S2½ values were much lower than that of normal polymers, indicating a small molecular volume of amylopectin, as a result of highly branched structure. Ignoring the difference of degree of branching, approximated dependences of 〈S2½ and intrinsic viscosity ([η]) on Mw for amylopectin in DMSO at 25°C were estimated to be 〈S2½ = 0.30Mw0.35 (nm) and [η] = 0.331Mw0.41 (mL g?1) in the Mw range studied. Moreover, from the 〈S2½ values of numberless fractions obtained from many experimental points in the SEC chromatogram detected with LS, the dependence of 〈S2½ on Mw for the II‐b sample was estimated also to be 〈S2½ = 0.34 Mw0.347, coinciding with the above results. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

3.
As a typical water-soluble polymer, ultra-high molecular weight (UHMW) partially hydrolyzed polyacrylamide (HPAM) has been widely used in various industries as thickeners or rheology modifiers. However, precise determination of its critical physical parameters such as molecular weight, radius of gyration (Rg) and hydrodynamic radius (Rh) were less documented due to their high viscosity in aqueous solution. In this work, the molecular structure of five UHMW-HPAM samples with different MW was elucidated by 1H and 13C NMR spectroscopy, and their solution properties were characterized by both static and dynamic light scattering. It is found that all the second virial coefficient (A2) values are positive and approaching zero, indicating of a good solvent of 0.5 M NaCl for UHMW-HPAM. The weight-average molecular weight (Mw) dependence of molecular size and intrinsic viscosity [η] for these series of HPAM polymers with MW ranging from 4.81 to 15.4 × 106 g·mol−1 can be correlated as Rg = 3.52 × 10−2Mw0.51, Rh = 1.97 × 10−2Mw0.51, and [η] = 6.98 × 10−4 Mw0.91, respectively. These results are helpful in understanding the relationship between molecular weight and coil size of HPAM polymers in solution, and offer references for quick estimation of molecular weight and screening of commercial UHMW-HPAM polymers for specific end-users.  相似文献   

4.
Three medium oil alkyd samples of different oil length [(I) 45%, (II) 50%, and (III) 55%] were synthesized with rubber seed oil. Dilute solution viscosity measurements were carried out on the alkyd in acetone and in toluene. The parameters investigated include intrinsic viscosity [η], Huggins constant (kH), and Mark‐Houwink Sakurada constants (κ and α). The [η] values for the alkyd samples were found to be larger in acetone than in toluene. The KH values showed a regular order in acetone but not in toluene. The KH values showed no regular order in their variation with the oil content of the alkyd samples and the solvent, but the values obtained are higher in acetone than in toluene. Correlation of molecular weight (M) with [η] was also examined. [η] was observed to increase with the increase in the molecular weight of the resins. The α values obtained are in reasonable agreement with the reported ranges of α values in good solvent. The characteristics examined suggest that acetone is a better solvent for rubber seed oil alkyd resin than toluene. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3073–3075, 2006  相似文献   

5.
Partially N‐acetylated chitosan was hydrolyzed by the cheap, commercially available, and efficient cellulase. The products, with different molecular weight, were comparatively investigated by GPC, FT‐IR, XRD, and NMR. The results show that the decrease of molecular weight led to transformation of crystal structure and increase of water‐solubility, but the chemical structures of residues were not modified. Superoxide anion radical and hydroxyl radical quenching assay were used for the evaluation of free radical scavenging activity of cellulase‐treated chitosan in vitro. Low molecular weight chitosan (LMWC3, Mw 1.7 × 103) exhibited high scavenging activity against free radical. It scavenged 79.3% superoxide radical at 0.1 mg mL?1. At 2.0 mg mL?1, scavenging percentage of initial chitiosan, LMWC1 (Mw 27.3 × 103), LMWC2 (Mw 5.9 × 103), and LMWC3 (Mw 1.7 × 103) against hydroxyl radical was 14.3%, 33.1%, 47.4%, and 65.9%, respectively. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

6.
The coil–globule transition for poly(methyl methacrylate) (PMMA) has been studied in a theta solvent, acetonitrile (Θ = 45 °C). The viscosity of PMMA was measured as a function of temperature in the range 26–55 °C. The contraction and expansion of the molecular chains are determined using the measured viscosity values. The temperature dependence of the intrinsic viscosity can be represented by a master curve in a versus |τ|M w1/2 (g1/2 mol−1/2) plot, where τ = |T − Θ|/T is the reduced temperature and Mw‐is the weight‐average molecular weight. A universal plot of reduced viscosity versus reduced blob parameter (N/Nc) shows the attainment of the collapsed state below the theta temperature. The dimensions of PMMA in acetonitrile (Mw = 3.15 × 106 g mol−1) decrease to 63 % at 26 °C of those in the unperturbed state. The results in this work are compared with those previously published. © 2000 Society of Chemical Industry  相似文献   

7.
Molecular, hydrodynamic, and thermodynamic properties of nylon 11 in meta cresol and 1,1,1,2,2,2-hexafluor-2-propanol (HFIP) solutions, as well as its distribution of molecular weights were investigated by means of viscosimetry, conductimetric titration of end groups, light scattering, and fractionation by successive precipitation. The studies were carried out on commercial samples as such (M n = 14,900, M w = 29,400) and on specimens prepared by solid-state postpolymerization of the former (M n = 35,000–43,000, M w = 91,000–104,000). The results show the expected normal or Flory–Schulz distribution of molecular weights on the commercial sample (U = 1.03), and a broadened distribution on the postpolymerized one (U = 1.42), in agreement with previous observations on solid-state postpolymerization of other polyamides. The intrinsic viscosity of the individual fractions was determined experimentally and the weight-average molecular weights were calculated from the data of the fractionation (number-average molecular weight and the mass fraction of polymer on each individual fraction) by means of an iterative numerical procedure. The parameters of the Mark–Houwink equation were, then, derived from the data of a large number of samples, including that corresponding to the whole, unfractionated polymers, spanning a range of about 100,000 units of molecular weight. The value of the exponent (a = 0.69) for solutions in meta cresol corresponds to the behavior of a linear, flexible macromolecule in a good-solvent medium. The solutions in HFIP employed for the light-scattering studies, on the other hand, display high values of the second virial coefficient (A2 = 7.8 × 10?3 ? 5.6 × 10?3 mol mL/g), suggesting that HFIP is a good solvent for nylon 11.  相似文献   

8.
The unperturbed dimensions and thermodynamic parameters of poly(vinylpyrrolidone) (PVP) have been studied in aqueous salt solutions, e.g. phosphates, mono- and dihydrogen phosphates, carbonates, sulphates of sodium and potassium. Values of K0 ( = [η]ΘM-1/2, where [η]Θ is intrinsic viscosity at the theta temperature and M is molecular weight) with Mw = 78 000 g mol-1 were found to range from 4·63×10-4 to 5·56×10-4 dl g-1, and root-mean-square end-to-end distances, 〈r201/2, ranging from 1·61×10-6 to 1·68×10-6cm were evaluated. Values of the characteristic ratio, Cn, the steric parameter, σ, and the enthalpy and the entropy of dilution parameters, χH and χS, have also been calculated, and the interaction parameter was found to be χ-0·5<-0·001 for aqueous salt solutions of PVP. ©1997 SCI  相似文献   

9.
The dilute solution properties of a cationic polyelectrolyte, poly(dimethyl sulfate quaternized dimethylaminoethyl methacrylate) [poly(DMAEM · C2H6SO4)], are studied by measurements of intrinsic viscosity, degree of binding, and flocculation application. The intrinsic viscosity of this polyelectrolyte is related to the type and concentration of added salt. The intrinsic viscosity behavior for cationic polyelectrolyte resulting from the electrostatic repulsive force of the polymer chain is contrasted with polyampholyte. The polyelectrolyte in the presence of KCl has a lower degree of binding, indicating that the proton ion (H+) is relatively difficult to bind to the CH3SO4? at the polymer end. The polymerization of DMAEM · C2H6SO4 in 0.5M KCl aqueous solution proceeded more easily than that of DMAEM · C2H6SO4 in pure water. The polymerization rate of DMAEM · C2H6SO4 is found to pass through an extreme value as a function of pH. Optimum flocculation, corresponding to the complete removal of turbidity in the supernatant, is achieved. Beyond the optimum flocculation, high polymer dosages redisperse the bentonite suspensions.  相似文献   

10.
A series of semi‐interpenetrating polymer networks (semi‐IPNs) films were prepared from 20 wt % of benzyl amylose (BA) of different Mw and castor oil‐based polyurethane (PU) in N,N‐dimethylformamide (DMF). The weight‐average molecular weight (Mw), and radii of gyration (<S2>z1/2) of benzyl amylose were determined by laser scattering measurement, and the results suggested BA was in a compact coil conformation in DMF. Furthermore, the properties and miscibility of the polyurethane/benzyl amylose (PUBA) films were studied by scanning electronic microscopy, differential scanning calorimetry, dynamic mechanical thermal analysis, ultraviolet–visible spectrophotometer, and tensile testing. The PUBA films possessed much higher optical transmittance and tensile strength than the pure PU film regardless of the molecular weight of BA, but lower values of elongation at break were observed. With decreasing of the BA Mw from 9.24 × 105 to 2.69 × 105, interestingly, the elongation at break of the films increased from 135 to 326%. This might be ascribed to the decrease of crosslinking density of PU networks and the enhancement in freedom of the molecular motion. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

11.
A novel polyamide has been successfully prepared through the reaction of 2,7‐dibromo‐9,9‐dioctylfluorene with 2,5‐dipiperazinedione in the presence of CuI, N,N′‐dimethylethylene diamine (DMEDA) and K2CO3 as base mixture and as catalyst. The structures of the monomer and the resulting model compound, as well as the structure, solution viscosity, solubility, molecular weights, thermal behavior, thermal stability, and light absorption and emission spectra of the resulting polyamide were characterized by means of FTIR, elemental analysis, 1H‐NMR, 13C‐NMR, DSC, TGA, GPC and UV–visible absorption, and fluorescence emission spectrophotometers. The polyamide possesses excellent solubility in organic solvents such as tetrahydrofuran (THF), N‐methyl‐2‐pyrrolidone (NMP), N,N‐dimethylacetamide (DMAc), N,N‐dimethylformamide (DMF), ethylacetate, acetone, ethanol, pyridine, chloroform, and toluene at room temperature. The polyamide had inherent viscosity of 0.65 dL/g, and molecular weights of Mn= 4.25 × 104 and Mw= 5.99 × 104 g/mol. The polyamide had glass transition temperature (Tg) of 138°C, and 10% weight loss at 350°C in nitrogen. The polyamide showed strong UV absorption and blue emission in solution and in solid state. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

12.
The refractive indices of poly(β-hydroxybutyric acid) (PHB) at four wavelengths have been determined via different procedures. Viscometric and light scattering measurements have been made on solutions of eight samples of PHB (Mw = 20·9 × 103?929 × 103 g mol?1) in 2,2,2-trifluoroethanol. From the dependences of intrinsic viscosity and of radius of gyration on molar mass, the conformation of PHB in dilute solution is shown to be that of a random coil. The findings are discussed in relation to existing conflicting evidence on the conformation of this polymer.  相似文献   

13.
To understand the molecular architectures of styrene‐butadiene four‐arm star (SBS) copolymers, a size exclusion chromatography combined with laser light scattering (SEC‐LLS) has been used to determine their weight‐average molecular weight (Mw) and radius of gyration (〈S21/2), and a new method for the establishment of the Mark‐Houwink equation from one sample has been developed. Based on the Flory viscosity theory, we successfully have reduced the 〈S21/2 values of numberless fractions estimated from many experimental points in the SEC chromatogram to intrinsic viscosities ([η]). For the first time, the dependences of 〈S21/2 and [η] on Mw for the four‐arm star SBS in tetrahydrofuran at 25°C were found, respectively, to be 〈S21/2 = 2.62 × 10?2 M (nm) and [η] = 3.68 × 10?2 M (mL/g) in the Mw range from 1.4 × 105 to 3.0 × 105. From data of [η] and 〈S21/2 for linear and star SBS, we have obtained the information about the branching, namely, the ratios (g and g′) of 〈S2〉 and [η] for star SBS to that of the linear SBS of the same molecular weight, which agree with theoretical predictions. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 96: 961–965, 2005  相似文献   

14.
The viscosity-average molecular weight, Mv, of a polymer is given operationally through its limiting viscosity number [η] and the Mark-Houwink equation [η] = KMvα, where K and α are empirical constants. If [η] is measured under different conditions, α and Mv will vary for the same sample. Mvα is the α-order moment about the origin of the differential weight distribution of the polymer. Practically, the results of a series of Mv measurements on the same polymer are equivalent to a cluster of fractional moments of the weight distribution, with orders between 0.55 and 0.80. It is shown that the first moment of this distribution, Mw, may be estimated reliably by a straightline plot of Mv against α-extrapolated to α equals 1. This simple expedient is effective although there are probably no molecular weight distributions in which the relation is strictly linear and there are no mathematical distributions for which the αth root of the αth moment is a linear function of α for all α. The deviation from linearity is small enough, however, that the real curve can be represented by a straight line over a short range of α. Thus, Mw can be measured accurately, but Mn, Mz, or the breadth of the distribution is not accessible by this method. Experimental and literature examples show that the precision of Mw estimated by this method compares well with that of primary methods for measuring this molecular weight average. If a linear relationship is observed with reliable α values, this appears to be a sufficient condition for estimation of a valid Mw.  相似文献   

15.
Static and dynamic light‐scattering techniques were used to study biodegradable thermoplastic poly(hydroxy ester ether) in N,N‐dimethylacetamide (DMAc). A weight‐average molecular weight MW = 6.4 × 104 g/mol, radius of gyration RG = 9.4 nm, second‐virial coefficient A2 = 1.05 × 10?3 mol mL/g2, translational diffusion coefficient D = 1.34 × 10?7 cm2/s, and hydrodynamic radius RH = 8.3 nm are reported. In addition, the effect of H2O on the polymer chain's conformation and architecture in a DMAc/H2O solution is evaluated. Results suggest that H2O makes the mixed solvent poorer as well as promotes polymer chain branching via intramolecular transesterification. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 1737–1745, 2001  相似文献   

16.
Summary Seven fractions of a (1→3)-α-D-glucan from Ganoderma lucidum have been studied by light scattering, sedimentation equilibrium, and viscometry in dimethylsulfoxide (DMSO) containing 0.25 M lithium chloride at 25°C. The intrinsic viscosity [η] - molecular weight relation for this glucan in the mixed solvent is found to be represented approximately by [η] = 0.071 M w 0.60 cm3 g−1 in the range of weight-average molecular weight M w studied, i.e., from 8 × 103 to 4.4 × 105. Its analysis based on current theories for wormlike chains shows that, without excluded volume effect, the (1→3)-α-D-glucan chain is characterized by a linear mass density of 380 nm−1, a Kuhn segment length of about 3 nm, and a diameter of 1.2 nm and is somewhat more extended but more flexible than amylose, a (1→4)-α-glucan, in DMSO. Received: 28 July 1998/Accepted: 18 August 1998  相似文献   

17.
A screening study of the solubility of poly(1,1-difluoroethylene) (PVF2, M?w = 2 × 105) at room temperature in a wide variety of dipolar aprotic species has facilitated the discovery of a series of new solvents (N-methyloxazolidone, cyclic-substituted ureas) and rationalization of the data in a two-dimensional solubility map involving their dipole moment, μ, and their hydrogenx-bond-accepting (HBA) power β (Taft solvatochromic parameter). This map may be used as a predictive tool for the research of new functional classes of solvents, such as N-substituted mixed amideester of phosphoric acid or N-substituted sulfurous diamides. The variations of the intrinsic viscosity of the polymer with solvent polarity may be quantitatively analyzed using a linear multiparametric correlation which emphasizes the two opposite contributions of cavitation effects (Hildebrand solubility parameter δ) and of polymer–solvent interactions (β) on the coil expansion: [η] (dL.g?1) = 0.792 - 1.2 × 10?3δ2(J.cm?3) + 1.59 β. Finally, 1,3-dimethyl-2-oxo-hexahydropyrimidine (N,N′-dimethylpropylene urea) leads to the highest value of the refractive index increment (dn/dc = ?0.065 mL.g?1 at λ = 632 nm), and thus appears as the best solvent for light-scattering measurements.  相似文献   

18.
To adopt a recently developed method for large scale fractionation (CPF = continuous polymer fractionation, a special kind of counter current extraction) to polyisobutylene (PIB), a systematic search for the best mixed solvent was performed. For this purpose, the essential parts of the phase diagrams solvent/nonsolvent/PIB were determined for 21 mixed solvents by cloud-point measurements; with eight systems of special interest, the molecular weight distributions of the polymers contained in the coexisting phases were also studied. On the basis of these experiments and of considerations concerning additional criteria for the performance of the continuous counter current extraction, the mixed solvent toluene/methyl ethyl ketone was chosen. First experiments with a PIB sample of Mw = 420,000 g/mol and a molecular nonuniformity U = (Mw/Mn) ? 1 of 2.3 yielded two high molecular weight fractions Mw = 1.1 × 106 and 0.6 × 106 resp., with U = 0.3 on a 100 g scale upon the application of four CPF steps.  相似文献   

19.
An in situ–generated tetrafunctional samarium enolate from the reduction of 1,1,1,1‐tetra(2‐bromoisobutyryloxymethyl)methane with divalent samarium complexes [Sm(PPh2)2 and SmI2] in tetrahydrofuran has proven to initiate the ring‐opening polymerization of ?‐caprolactone (CL) giving star‐shaped aliphatic polyesters. The polymerization proceeded with quantitative conversions at room temperature in 2 h and exhibited good controllability of the molecular weight of polymer. The resulting four‐armed poly(?‐caprolactone) (PCL) was fractionated, and the dilute‐solution properties of the fractions were studied in tetrahydrofuran and toluene at 30°C. The Mark–Houwink relations for these solvents were [η] = 2.73 × 10?2Mw0.74 and [η] = 1.97 × 10?2Mw0.75, respectively. In addition, the unperturbed dimensions of the star‐shaped PCL systems were also evaluated, and a significant solvent effect was observed. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 175–182, 2006  相似文献   

20.
The radiation-induced homopolymerization of acrylonitrile (AN) in dimethylformamide (DMF) has been followed in detail over a range of monomer concentrations. In all cases a short induction period was observed which was equivalent to 1.6 × 104 rad. The initial rates of polymerization for solutions of AN in DMF of mole fraction (×) of 0.33, 0.20, and 0.10 are 1.44 × 10?5, 7.23 × 10?6, and 2.79 × 10?6 mole/dm3rad, respectively. Deviations of the polymerization pathway from the standard unity in monomer dependence are examined in terms of radical production ratios to the monomer and the solvent (ΦMS) and the polymer together with the solvent (Φps), for various mole fractions of AN in DMF. Thus, an indirect route to Gradical(events) is provided together with the corresponding k/kt values.  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