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 15 毫秒
1.
Linear and branched polyethylene were fractionated by solvent gradient chromatography. The fractions were characterized by light scattering and viscosity measurements. The relationship between the molecular weight and the limiting viscosity number and that between the molecular weight and the radius of gyration in α-chloronaphthaline at 127°C were determined. The exponent of the equation between the ratio 〈gη〉Θ of the limiting viscosity numbers of branched and linear fractions of the same molecular weight Mw and the corresponding ratio 〈gηΘ,w of the radii of gyration for Θ-conditions was By this equation, the long chain branching of high pressure polyethylene can be determined from 〈gη〉, the ratio of the limiting viscosity numbers in thermodynamically good solvents, which is within the experimental limits of error identical to 〈gηΘ in a Θ-solvent. Almost independent of the molecular weight, for the branched polyethylene fractions 1,4 long chain branching points per 1000 C-atoms were obtained.  相似文献   

2.
Summary: The oligomeric A2 plus monomeric B3 synthetic methodology provided highly branched, poly(ether urethane)s based on TMP (B3) and isocyanate endcapped polyethers. 13C NMR spectroscopic assignments for the branched polyurethanes were verified using model urethane‐containing compounds based on TMP and a monofunctional isocyanate (either cyclohexyl or phenyl isocyanate (PI)). Derivatization of hydroxyl endgroups with trifluoroacetic anhydride enhanced the 13C NMR resolution in spectra for branched polyurethanes. The 13C NMR resonance for the linear unit exhibited a broad shoulder due to quaternary carbons that were attributed to cyclic species in the highly branched polyurethanes. The classical DB calculation revealed the efficiency of the B3 monomer for branching; however, an equation that incorporated the linear contribution of the A2 oligomer provided a more accurate DB for highly branched polyurethanes.

SEC chromatograms for increasing addition of A2.  相似文献   


3.
Optical rotatory dispersion (ORD) and circular dichroism (CD) of poly(L -lactate) were investigated and were compared with those of optically active oligo(L -lactate), prepared by the stepwise method. Two-term DRUDE equation was applied to analyse the ORD curves using n-π* and π-π* transition bands as the critical wavelengths, Λ0 and Λ1. No drastic difference in ORD between poly(L -lactate) and oligo(L -lactate) was detected in chloroform and dioxane, which suggests that the polymer does not form any special conformation in these solvents. However, in acetonitrile it was found that the value of ?A1/A0 (A1 and A0 are constants characteristic of π-π* and n-π* transitions of ester chromophores, respectively) of the polymer differed from that of the oligomer (n = 3).  相似文献   

4.
For number‐average molecular weight (M n) below 1 × 104 g mol?1, the comparison of cold crystallization temperature and spherulite growth rate and crystallinity of linear 1‐arm, 2‐arm, and branched 4‐arm poly(L ‐lactide)/poly(D ‐lactide) blends exhibits that the effects of chain directional change and branching significantly disturb stereocomplex crystallization. In contrast, the comparison of glass transition and melting temperatures of linear 1‐arm, 2‐arm, and branched 4‐arm poly(L ‐lactide)/poly(D ‐lactide) blends indicates that the effects of chain directional change and branching insignificantly alter and largely increase the segmental mobility of the blends, respectively, and the crystalline thickness of the blends is determined by M n per one arm not by M n and is not affected by the molecular architecture.

  相似文献   


5.
The tacticity of poly(α-methylstyrene) was measured by 1H NMR spectroscopy. The polymerisation initiated in tetrahydrofuran by butyllithium gives a dependence on the initial monomer concentration and a S-shaped plot of log(m/r) vs. 1/T, which is not influenced by a stronger reaction medium like 1,2-dimethoxyethane or hexamethylphosphoric triamid or by addition of LiB(C6H5)4. The activation parameters of the two straight lines A and B limiting the experimental curve are according to the equation The temperature dependence of the propagation rate constant kw measured dilatometricly between ?11°C and ?75°C gives a linear Arrhenius plot. The rate constant of depolymerisation kd was calculated using the equilibrium monomer concentration. The activation parameters are The influence of ionic dissociation on the tacticity is discussed. The results lead to the conclusion, that the contact ion pair, the solvent separated ion pair, and the free anion generate tacticities, which are undistinguishable by the applied experimental methods.  相似文献   

6.
The Huggins coefficient kH of solutions of statistical styrene/butadiene copolymers in cyclohexane (at 25 °C) and 2-pentanone (theta temperature 21 °C) is discussed. With an increase in branching the values of (i. e. kH at the theta temperature) increase from 0,5 to 0,73. Lower values in cyclohexane (good solvent) correlate with the viscosity expansion factor αη3, and the dependence is well fitted by the equation where is a function of the degree of branching, whereas C0 only slightly exceeds the value found for linear polymers.  相似文献   

7.
The intrinsic viscosity, osmotic pressure and light scattering of solutions of fractionated samples of low density polyethylene (LDPE) were measured to study its long chain branching. The following relation was obtained between the two branching indices: gη is the ratio of the intrinsic viscosities of the branched and linear polymers with the same molecular weight, and gs is the corresponding ratio of the radii of gyration. The average number of long branches per molecule, nw, was calculated according to the theory of ZIMM and STOCKMAYER derived for random branching, and it was found that the dependence of nw on the molecular weight is expressed by with α = 0.85. This relation indicates that the long branching frequency or density decreases gradually with molecular weight for LDPE.  相似文献   

8.
Summary: The oligomerization of γ‐branched α‐olefins in the presence of catalytic systems based on group‐4 metallocenes with C2v symmetry has been investigated. The highest reactivity was obtained by using dimethylsilyl‐bis(cyclopentadienyl)zirconium‐dichloride activated by methylaluminoxane. Highly regioregular dimers were selectively obtained for hindered γ‐branched monomers, while the less hindered ones produced higher molecular weight oligomers. A molecular modeling approach was used to rationalize the experimental results. In fact, a decrease in the β‐hydrogen elimination barrier and an increase in the insertion barrier with the monomer bulkiness were calculated.

General structure of the obtained dimers.  相似文献   


9.
Summary: Controlled polymerization of styrene in toluene was achieved by atom transfer radical polymerization (ATRP) using hyperbranched polyglycidol‐supported multidentate amine ligands/CuIBr catalyst systems. These catalyst systems with nanoscopic dimensions were more active than the corresponding low‐molecular‐weight ligands. The controlled/living nature of the polymerization is supported by linear first order plots (ln[M]o/[M] versus time) and a linear increase of versus conversion as well as low polydispersity. Similar controlled polymerization of methyl methacrylate was possible in acetonitrile. Up to 97% of total copper used for polymerization could be removed from the polymer by simple precipitation in methanol and filtration.

Molecular weight characteristics for styrene polymerization using PG‐triamine as a ligand.  相似文献   


10.
A series of oligourethan dimers having asymmetric carbon atoms in the main chain was prepared by the reaction between 1-alkyl-2-acetoxyethyl isocyanates and methyl l-alkyl-2-hydroxyethylcarbamates; their melting points, rotatory powers and RF-values were determined. The dimers prepared are represented by a general formula, where R and R' are H, methyl (L and DL-forms), i-propyl (L and DL), i-butyl (L and DL) or benzyl (L) groups. Certain relationships have been found between the structure and rotatory power resp. RF-value.  相似文献   

11.
Here, the synthesis, characterization, and photovoltaic properties of four new donor–acceptor copolymers are reported. These copolymers are based on 4,4‐difluoro‐cyclopenta[2,1‐b:3,4‐b′] dithiophene as an acceptor unit and various donor moieties: 4,4‐dialkyl derivatives of 4H‐cyclopenta[2,1‐b:3,4‐b′]dithiophene and its silicon analog, dithieno[3,2‐b:2′,3′‐d]‐silol. These copolymers have an almost identical bandgap of 1.7 eV and have a HOMO energy level that varies from ?5.34 to ?5.73 eV. DSC and X‐ray diffraction (XRD) investigations reveal that linear octyl substituents promote the formation of ordered layered structures, while branched 2‐ethylhexyl substituents lead to amorphous materials. Polymer solar cells based on these copolymers as donor and PC61BM as acceptor components yield a power conversion efficiency of 2.4%.

  相似文献   


12.
This article reports the synthesis of novel amphiphilic triblock copolymers with a semi‐branched PLURONIC®7R structure by atom transfer radical polymerization (ATRP) in aqueous media. Poly(ethylene oxide)s (PEOs) with molecular weights 10 000 and 16 000 were end‐functionalized and used as bifunctional macroinitiators for the polymerization of oligo(propylene oxide) monomethacrylate by ATRP in a 1/3 v/v water/methanol mixture and in a 1/1 v/v water/1‐propanol mixture. Deviations from first‐order kinetics with respect to the monomer concentration were observed indicating that termination reactions were taking place. However, linear plots were obtained, when ln[M]0/[M] was plotted against time2/3 as suggested by Fischer. The effect on the control of the polymerization by adding Cu(II)Br2 to the polymerization medium has been investigated. When 10 mol‐% of Cu(II)Br2 was substituted for Cu(I)Br, normal first‐order kinetics were observed. A large reduction in the rate of polymerization was observed for the polymerization initiated by bifunctional PEO10 000 initiator, but almost no reduction in the rate of polymerization was observed, when the bifunctional PEO16 000 initiator was used. When the polymerizations were conducted in 1/1 v/v water/1‐propanol, unexpectedly high rates of polymerization were observed.

Synthesis of amphiphilic block copolymers with a semi‐branched PLURONIC®R architecture using ATRP.  相似文献   


13.
A method for the evaluation of the ratio of the mean square inertia radii \documentclass{article}\pagestyle{empty}\begin{document}$ (\overline {R^2 } )^{{1 \mathord{\left/ {\vphantom {1 2}} \right. \kern‐\nulldelimiterspace} 2}} $\end{document} of branched (b) and linear (1) molecules \documentclass{article}\pagestyle{empty}\begin{document}$ g[\eta ] = {{(\overline {R_{\rm b}^2 } )_{[\eta ]} } \mathord{\left/ {\vphantom {{(\overline {R_{\rm b}^2 } )_{[\eta ]} } {(\overline {R_{\rm l}^{\rm 2} } }}} \right. \kern‐\nulldelimiterspace} {(\overline {R_{\rm l}^{\rm 2} } }})_{[\eta ]} $\end{document} from intrinsic viscosity measurements is proposed. The evaluation is based on the solution of the equation: where erf \documentclass{article}\pagestyle{empty}\begin{document}$ x = \frac{2}{{\sqrt \pi }}\int\limits_0^{\rm x} {{\rm e}^{ ‐ t^2 } } {\rm d}t $\end{document}; a is the exponent in the Mark-Kuhn-Houwink relation [η1] = KMa, ε = (2a?1)/3 and ?(a) is the shielding function, tabulated by Debye and Bueche. The method is used for the evaluation of g[η] in Θ- and in good solvents from experimental data published in the literature. The results show that it is possible to find the dimensions of branched molecules from intrinsic viscosity measurements in good solvents using the above formula. These results are obtained by the analysis of changes of the macromolecular dimensions in different solvents and in part by the results of second virial coefficient measurements of different polymers.  相似文献   

14.
In order to assign all ? CH2Cl resonances observed by 1H NMR in poly(vinyl chloride), (PVC), a preliminary study of the diastereoisomers of molecular models for the groups CH2Cl? CHCl? [II] and is used, in conjunction with spin echo experiments. The complex region [3,5–4,0 ppm] which corresponds to saturated ? CH2Cl chloromethyl protons is completely assigned. A separate evaluation of ? CHCl? CH2Cl [II], ? CH2? CH2Cl [III] and is proposed and tested on raw PVC samples. The main interest of this 1H NMR determination is the fact that raw PVC can be used and that it gives the ratio [II]/[III] which is not accessible by 13C NMR on reduced samples.  相似文献   

15.
A novel series of aluminum complexes supported by asymmetric [ONNO′]‐type Salan ligands ( 1–4 ) is successfully synthesized. Polylactides (PLAs) with molecular weights close to the theoretical values and narrow polydispersity indices (PDIs) can be obtained. Kinetic studies reveal first‐order polymerization kinetics and a linear relationship between the molecular weight and the percentage conversion. Complexes 1–4 display heteroselective polymerization to give heterotactic‐enriched PLAs with Pr values ranging from 0.64 to 0.74. The character of the ligand substituents from the two different phenolic rings and the complex fluxionality arising from the ligand interconversion are suggested to be responsible for the different levels of stereoselectivity and activity.

  相似文献   


16.
Summary: Branched poly(arylene ether)s were prepared in an oligomeric A2 + B3 polymerization of phenol endcapped telechelic poly(arylene ether sulfone) oligomers as A2 and TFPPO as trifunctional monomer B3. The molar mass of the A2 oligomer significantly influenced the onset of gelation and the DB. A high level of cyclization during polymerization of low molar mass A2 oligomers (U3 = 660 and U6 = 1 200 g · mol?1) led to a high conversion of functional groups in the absence of gelation, and the level of cyclization reactions in the polymerization decreased as the molar mass of the A2 oligomer was increased. The pronounced steric effect in the polymerization of higher molar mass A2 oligomers (U8 = 1 800 and U16 = 3 400 g · mol?1) resulted in low reactivity of the third aryl fluoride in the B3 monomer. As a result, only slightly branched (U8 = 1 800 g · mol?1) or nearly linear (U16 = 3 400 g · mol?1) high molar mass products were obtained with higher molar mass A2 oligomers. The branched polymers exhibited lower Mark‐Houwink exponents and [η] relative to linear analogs, and differences between the branched polymers and linear analogs were less significant as the molar mass of the A2 oligomers was increased due to a decrease in the overall DB.

  相似文献   


17.
Analysis of the 1H NMR spectra of \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm CH}_{\rm 3} \mathop {\rm C}\limits^ \oplus {\rm OSbF}_{\rm 6}^ \ominus $\end{document} /β-propiolactone and of the model systems \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm CH}_{\rm 3} \mathop {\rm C}\limits^ \oplus {\rm OSbF}_{\rm 6}^ \ominus $\end{document}/(CH3)2O and \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm CH}_{\rm 3} \mathop {\rm C}\limits^ \oplus {\rm OSbF}_{\rm 6}^ \ominus $\end{document}/CH3COOCH3 in liquid SO2 in the temperature region ?70 to ?20°C revealed that the reaction of the acetylium cation with β-propiolactone leads to the cyclic six-membered oxonium ion 4 , participating as an intermediate in the initial stage of polymerization.  相似文献   

18.
Aliphatic AB2 functional polyesters were conveniently prepared by the ring opening polymerization of ε‐caprolactone and L ‐lactide in the presence of the AB2 functional initiator 2,2‐bis(hydroxymethyl)propionic acid (bis‐MPA) and Sn(Oct)2 as the catalyst. In L ‐lactide polymerization, both bis‐MPA hydroxyl groups initiated the polymerization reaction, but for ε‐caprolactone polymerization this depended on the monomer to initiator to catalyst ratio. Initiation at two hydroxyl groups occurred at high monomer to initiator ([M]/[I]) ratio and at high Sn(Oct)2 to monomer ratio. The melting temperatures of the AB2‐functional PLLA and PCL polymers were comparable to linear polymers with a equal to the per arm in the polymer.

  相似文献   


19.
An improved synthesis of photosensitive homopolymers containing aryltriazene chromophores covalently incorporated into the polymer backbone is reported. Such photopolymers proved to have promising properties for novel UV‐laser applications. A homologous series of new aryltriazene polymers with increasingly branched side chains (R = Me, Et, iPr, tBu) was synthesized and characterized. Homogeneous thin films with thicknesses from ≈15 to >150 nm were prepared by spin‐coating. Photodecomposition was studied in solution and on thin films. Polymers with increasingly branched and bulky substituents showed decreasing photodissociation rates. NMR studies suggested an enhanced hindrance of the N(2)–N(3) bond rotation in the aryltriazene moiety with increasing steric demand of the substituents.

  相似文献   


20.
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