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.
To investigate the effects of substituents attached to the silicon atom on the thermal rearrangement reactions of α‐silyl alcohols, the thermal rearrangement reactions of dimethylsilyl methanol (CH3)2SiHCH2OH and vinylsilyl methanol CH2?CHSiH2CH2OH were studied by ab initio calculations at the G3 level. Geometries of various stationary points were fully optimized at the MP2(full)/6‐31G(d) and MP2(full)/6‐311G(d,p) levels, and harmonic vibrational frequencies were calculated at the same levels. The reaction paths were investigated and confirmed by intrinsic reaction coordinate (IRC) calculations at the MP2(full)/6‐31G(d) level. The results show that two dyotropic reactions could occur when (CH3)2SiHCH2OH or CH2?CHSiH2CH2OH is heated. One is Brook rearrangement reaction (reaction A), and the dimethylsilyl or vinylsilyl groups migrates from carbon atom to oxygen atom coupled with a simultaneous migration of a hydrogen atom from oxygen atom to carbon atom passing through a double three‐membered ring transition state, forming dimethylmethoxylsilane (CH3)2SiHOCH3 or methoxylvinylsilane CH2?CHSiH2OCH3; the other is a hydroxyl group migration (reaction B) from carbon atom to silicon atom, coupled with a simultaneous migration of a hydrogen atom from silicon atom to carbon atom, via a double three‐membered ring transition state, forming trimethylsilanol (CH3)3SiOH or methylvinylsilanol CH3SiH(OH)CH?CH2. The G3 barriers of the reactions A and B were computed to be 312.8 and 241.4 kJ/mol for (CH3)2SiHCH2OH, and 317.6 and 233.7 kJ/mol for CH2?CHSiH2CH2OH, respectively. On the basis of the MP2(full)/6‐31G(d) optimized parameters, vibrational frequencies, and G3 energies, the reaction rate constants k(T) and equilibrium constants K(T) were calculated using canonical variational transition state theory (CVT) with centrifugal‐dominant small‐curvature tunneling (SCT) approximation over a temperature range of 400–1800 K. The influences of methyl and vinyl groups attached to the silicon atom on reactions are discussed. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

2.
Water‐soluble palladium complexes cis‐[Pd(L)(OAc)2] ( 1–8 ) (L represents a diphosphine ligands of the general formula CH2(CH2PR2)2, where for a : R ? (CH2)6OH; b–g : R ? (CH2)nP(O)(OEt)2, n = 2–6 and n = 8; h : R ? (CH2)3NH2) have been employed, after activation with a large excess of HBF4, for emulsion polymerization of alkenes (propene, butene, and their equimolar mixtures) with carbon monoxide. Aliphatic polyketone lattices with a high solid content (21%), high molecular weight (6.3 × 104 g mol?1), and narrow polydispersities (Mw/Mn ≈ 2) were isolated. The catalytic activity of the dicationic palladium (II) based catalysts, C1–C8 is highly dependent on the length of the alkyl chain of the ligand. Catalyst 3 proved to be highly active for propene/CO copolymers, whereas 6 is active for butene/CO and propene/CO‐butene/CO systems. The presence of methyl β‐cyclodextrin, as a phase‐transfer agent, and undecenoic acid, as an emulsifier, increase the molar mass and the stability of the polyketones and finally the activity of the catalyst. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 6715–6725, 2009  相似文献   

3.
The spreading behaviour of binary and ternary equimolar mixtures of siloxane surfactants of general formula [(CH3)3SiO]2CH3Si(CH2)3 (OCH2CH2) nOCH3, n = 3–9, has been investigated. The mixtures show a pronounced temperature dependence on the initial spreading rate. Mixtures imitating the average oligoethylene glycol chain length n = 5 are the fastest spreaders at 15 °C. At 23 °C and 40 °C these mixtures spread fastest sucking n = 6 and n = 8, respectively. For a given average chain length an increasing length difference between the components of the binary mixtures reduces the initial spreading rate. Nevertheless, substantial differences between the phase transition temperature Tc from the lamellar phase (Lα) into the two‐phase state (2Φ) and the actual spreading temperature are tolerated. A clear relation between phase transition temperature Tc and initial spreading rate does not exist. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

4.
A new series of thermotropic liquid‐crystalline (LC) polyesters were prepared from a diacyl chloride derivative of 4,4′‐(terephthaloyldioxy)‐di‐4‐phenylpropionic acid (PTP) and glycols with a different number of methylene groups (n) [HO(CH2)n OH, n = 6–10, 12] by high‐temperature solution polycondensation in diphenyl oxide. PTP6/10 and PTP6/hydroquinone (H) LC copolyesters were also prepared according to a similar procedure. The chemical structure, LC, phase‐transition behaviors, thermal stability, and solubility were characterized by elemental analysis, Fourier transform infrared spectroscopy, 1H and 13C NMR spectra, differential scanning calorimetry (DSC), thermogravimetric analysis, and a polarizing light microscope. The melting and isotropization temperatures decreased in a zigzag manner as the number of n increased. All of the polyesters formed a nematic phase with the exception of PTP8. The temperature ranges of the mesophase (ΔT) were much wider for the polyesters with an odd number of n's than those with an even number. ΔT increased markedly for the PTP6/10 and PTP6/H copolyesters. The in vitro degradations of the polymers were ascertained by enzymatic hydrolysis and alkaline hydrolysis. The model compound, PTP dihexylester, was synthesized and found to be degraded into terephthalic acid, 3‐(4‐hydroxyphenyl)propionic acid, and 1‐hexanol by Rhizopus delemar lipase, but PTPn homopolyesters and PTP6/10 and PTP6/H copolyesters were resistant to Rhizopus delemar hydrolysis. They were degradable in a sodium hydroxide buffer solution of pH 12 at 60 °C, depending on the number of n's and the copolymer composition. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 3043–3051, 2001  相似文献   

5.
《Polyhedron》1988,7(6):449-462
The complexes [ML*(NO)Cl(OR)] {L* = HB(3,5-Me2C3HN2)3; M= Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2, 5, 6; M = W, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; CH2(CF2)3CH2OH; CHMeCH2CMe2OH} and [ML*(NO)(OR)2] {M = Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; M = W,R = CH2CH2X, X= Cl, OMe or OEt; (CH2)nOH, n = 2,4–6; CH2(CF2)3CH2OH} have been prepared from [ML*(NO)Cl2] and the appropriate alcohol in the presence of NEt3 or NaCO3, and have been characterized by IR, 1H NMR and mass spectroscopy.  相似文献   

6.
Hydrogen exchange reactions between lithium and sodium compounds, MX (M=Li: X=H, CH3, NH2, OH, F; M=Na: X=CH3), and the corresponding hydrides, HX, have been modelled by means of ab initio calculations including electron correlation and zero point energy (ZPE) corrections. Small or no activation barriers (from the initial complexes) are encountered in systems involving lone pairs (10.8, 2.4, 0.0 kcal/mol for X=NH2, OH, F, respectively). Since the association energies of the initial complexes are much larger (21.0, 20.4, 23.5 kcal/mol, respectively; MP2/6–31+G*/6–31+G* + ZPE), such exchange reactions should occur spontaneously in the gas phase. The methyl systems (X=CH3) have the largest barriers: 26.7 (M=Li) and 31.7 (M=Na) kcal/mol (MP2/6–31+G*/6–31G* + ZPE), and the initial complexes are only weakly bound. The significance of these systems as models for hydrogen exchange reactions in complexes of electropositive transition metals is discussed. However, the gegenion-free exchange of hydrogen between CH3 and CH4 has a much lower, 11.8 kcal/mol barrier (MP2/6–31+G*/6–31+G* + ZPE). All the transition structures are highly ionic (charges on the metals > +0.8). The effect of aggregation has been considered by examining the hydrogen exchange between (LiX)2 and HX(X=H, CH3, NH2, OH). Although these dimer reactions formally involve six, instead of four electrons, no “aromatic” preference is observed.  相似文献   

7.
Equimolar reactions of Ph n M(OPr i )2 (where M?=?As and Sb) with Schiff bases [OHC6H4CH=N(R)OH] in benzene solution yield organoarsenic and -antimony derivatives, (where M?=?As and Sb; n?=?1 and 3; R?=?–CH2CH(CH3)–, –(CH2)3–, –(CH2)2–, and –C(CH3)2CH2–). All these derivatives have been characterized by elemental analyses and molecular weight measurements, and structures have been proposed on the basis of IR, NMR (1H and 13C), and FAB-mass studies. Schiff bases and their corresponding organoantimony derivatives have been screened for antimicrobial activity against Aspergillus flavus (fungus) and Escherichia coli (bacteria).  相似文献   

8.
The rate constant of the title reaction is determined during thermal decomposition of di-n-pentyl peroxide C5H11O( )OC5H11 in oxygen over the temperature range 463–523 K. The pyrolysis of di-n-pentyl peroxide in O2/N2 mixtures is studied at atmospheric pressure in passivated quartz vessels. The reaction products are sampled through a micro-probe, collected on a liquid-nitrogen trap and solubilized in liquid acetonitrile. Analysis of the main compound, peroxide C5H10O3, was carried out by GC/MS, GC/MS/MS [electron impact EI and NH3 chemical ionization CI conditions]. After micro-preparative GC separation of this peroxide, the structure of two cyclic isomers (3S*,6S*)3α-hydroxy-6-methyl-1,2-dioxane and (3R*,6S*)3α-hydroxy-6-methyl-1,2-dioxane was determined from 1H NMR spectra. The hydroperoxy-pentanal OHC( )(CH2)2( )CH(OOH)( )CH3 is formed in the gas phase and is in equilibrium with these two cyclic epimers, which are predominant in the liquid phase at room temperature. This peroxide is produced by successive reactions of the n-pentoxy radical: a first one generates the CH3C·H(CH2)3OH radical which reacts with O2 to form CH3CH(OO·)(CH2)3OH; this hydroxyperoxy radical isomerizes and forms the hydroperoxy HOC·H(CH2)2CH(OOH)CH3 radical. This last species leads to the pentanal-hydroperoxide (also called oxo-hydroperoxide, or carbonyl-hydroperoxide, or hydroperoxypentanal), by the reaction HOC·H(CH2)2CH(OOH)CH3+O2→O()CH(CH2)2CH(OOH)CH3+HO2. The isomerization rate constant HOCH2CH2CH2CH(OO·)CH3→HOC·HCH2CH2CH(OOH)CH3 (k3) has been determined by comparison to the competing well-known reaction RO2+NO→RO+NO2 (k7). By adding small amounts of NO (0–1.6×1015 molecules cm−3) to the di-n-pentyl peroxide/O2/N2 mixtures, the pentanal-hydroperoxide concentration was decreased, due to the consumption of RO2 radicals by reaction (7). The pentanal-hydroperoxide concentration was measured vs. NO concentration at ten temperatures (463–523 K). The isomerization rate constant involving the H atoms of the CH2( )OH group was deduced: or per H atom: The comparison of this rate constant to thermokinetics estimations leads to the conclusion that the strain energy barrier of a seven-member ring transition state is low and near that of a six-member ring. Intramolecular hydroperoxy isomerization reactions produce carbonyl-hydroperoxides which (through atmospheric decomposition) increase concentration of radicals and consequently increase atmospheric pollution, especially tropospheric ozone, during summer anticyclonic periods. Therefore, hydrocarbons used in summer should contain only short chains (<C4) hydrocarbons or totally branched hydrocarbons, for which isomerization reactions are unlikely. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 875–887, 1998  相似文献   

9.
Reactions that proceed within mixed ethylene–methanol cluster ions were studied using an electron impact time-of-flight mass spectrometer. The ion abundance ratio, [(C2H4)n(CH3OH)mH+]/[(C2H4)n(CH3OH)m+], shows a propensity to increase as the ethylene/methanol mixing ratio increases, indicating that the proton is preferentially bound to a methanol molecule in the heterocluster ions. The results from isotope-labelling experiments indicate that the effective formation of a protonated heterocluster is responsible for ethylene molecules in the clusters. The observed (C2H4)n(CH3OH)m+ and (C2H4)n(CH3OH)m–1CH3O+ ions are interpreted as a consequence of the ion–neutral complex and intracluster ion–molecule reaction, respectively. Experimental evidence for the stable configurations of heterocluster species is found from the distinct abundance distributions of these ions and also from the observation of fragment peaks in the mass spectra. Investigations on the relative cluster ion distribution under various conditions suggest that (C2H4)n(CH3OH)mH+ ions with n + m ≤ 3 have particularly stable structures. The result is understood on the basis of ion–molecule condensation reactions, leading to the formation of fragment ions, $ {\rm CH}_2=\!=\mathop {\rm O}\limits^ + {\rm CH}_3 $ and (CH3OH)H3O+, and the effective stabilization by a polar molecule. The reaction energies of proposed mechanisms are presented for (C2H4)n(CH3OH)mH+(n + m ≤ 3) using semi-empirical molecular orbital calculations.  相似文献   

10.
Six new coordination complexes, [Cd(η 2-OOCCH=(CH3)CFc)2(bix)]2·(CH3OH)0.5 (1), [Zn(η 2-OOCCH=(CH3)CFc)(η 1-OOCCH=(CH3)CFc)(bix)]2·(H2O)0.5 (2), [Zn(η 2-OOCCH=(CH3)CFc)2(pbbm)]2·(CH3OH)2 (3), {[Mn(η 1-OOCCH=(CH3)CFc)2(bbbm)(H2O)2]·(CH3OH)3}n (4), {[Cd(η 1-OOCCH=(CH3)CFc)2(bbbm)]·(CH3OH)2}n (5), and [Cd(η 2-OOCCH=(CH3)CFc)2(pmbbm)]n (6) {Fc?=?(η 5-C5H4)Fe(η 5-C5H4), bix?=?1,4[bis(imidazol-1-ylmethyl)benzene], pbbm?=?1,1′-[(1,4-propanediyl)bis-1H-benzimidazole], bbbm?=?1,1′-[(1,4-butanediyl)bis-1H-benzimidazole)], pmbbm?=?1,1′-[(1,4-pentanediyl)bis-1H-benzimidazole]}, were prepared and characterized. X-ray crystallographic analysis reveals that 1–3 are dimers bridged by bix and pbbm. Complexes 4–6 are one-dimensional (1-D) structures bridged by bbbm and pmbbm, respectively. Various ππ interactions were discovered in 1–6 that make significant contributions to molecular self-assembly. Solution differential pulse voltammetry of 1–6 indicates that the half-wave potentials of the ferrocenyl moieties in these complexes shift to positive potential compared with that of 3-ferrocenyl-2-crotonic acid.  相似文献   

11.
在本文中,提出了极性基团电子相关能贡献的定义,并在MP2-OPT2/6-311++G(d)水平上计算了CH3(CH2) mOH( m=0-4)体系中HO-、CH3-和-CH2-基团电子相关能贡献值。计算结果表明,在CH3(CH2) mOH( m=0-4)体系中端基HO-、CH3-基团电子相关能贡献值 Ecorr(HO-)和 Ecorr(cH3-)的数值随着 m的增加而逐渐减小。同一体系中a -CH2-基团电子相关能贡献值大于其它-CH2-基团电子相关能贡献值,在CH3(CH2) mOH( m=1-4)体系中,距离端基HO-基团越远的-CH2-基团其电子相关能贡献值越小;通过计算结果可以推断,在CH3(CH2) mOH体系中随着 m的逐渐增加,相对远离端基HO-的-CH2-基团的电子相关能贡献值表现出收敛趋向并将趋于不变,此-CH2-基团可看作一个标准的亚甲基而且其 Ecorr(-cH2-)的数值在CH3(CH2) mOH体系中具有传递性。在MP2-OPT2/6-311++G(d)水平上对CH3(CH2) mOH( m=2-4)体系的计算结果和应用Gaussian 98程序在MP2/6-311++G(d)//HF/6-311++G(d)水平上对CH3(CH2) mOH( m=2-7)体系的计算结果均表明,体系总电子相关能与( m-1)呈 中 m是体系中亚甲乙烯基的数值。  相似文献   

12.
Reactions of the isopropoxides of some of the lighter lanthanons with bidentate -ketoimines, such asAAH-n-C4H9 andAAH-C6H5 (donor system: N,OH) and tridentate -ketoimines such asAA(CH2CH2)H2 andAA(CH2CHCH3)H2 (donor system: HO,N.OH) have led to products of the typesLn(O-i-C3H7)3n (AA-R) n ,Ln(Oi-C3H7) (AAR') andLn 2(AAR')3 [Ln=La(III), Pr(III) or Nd(III);n=1 or 2;R=-n-C4H9 or-C6H5 andR'=-CH2CH2-or-CH2CHCH3-]. Some undergo exchange reactions with an excess oftert-butanol, leading to the corresponding complexesLn(O-tert-C4H9)3n (AA-n-C4H9) n andLn(O-tert-C4H9) (AA-CH2CH2). All these have been characterised by elemental analysis, molecular weight determinations and their ir spectra. A thermogravimetric analysis of the diisopropoxy derivatives has also been carried out.
Schiff-Basen Derivate von Lanthaniden-Synthese von La(III), Pr(III) und Nd(III) chelaten mit -Ketoiminen
Zusammenfassung Reaktionen von Lanthanid-Isopropoxiden mit zweizähnigen -Ketoiminen [AAH-n-C4H9 undAAH-C6H5; Donorsystem: N,OH] und dreizähnigen -Ketoiminen [AA(CH2CH2)H2 undAA(CH2CHCH3)H2; Donorsystem: OH, N,OH] führten zu Produkten vom, TypLn(O-i-C3H7)3-n (AA-R) n ,Ln(O-i-C3H7) (AAR') undLn 2(AAR')3 [Ln=La(III), Pr(III) oder Nd(III);n=1 oder 2;R=n-C4H9 oder C6H5 undR'=CH2CH2 oder CH2CHCH3]. Einige Komplexe unterliegen bei Behandlung mit einem Überschuß vontert-Butanol einer Austauschreaktion, die zu den entsprechenden Butoxid-Komplexen führt [Ln(O-tert-C4H9)3-n , (AA-n-C4H9) n undLn(O-tert-C4H9) (AACH2CH2)]. Alle Derivate wurden mittels Elementaranalyse, Molgewichtsbestimmung und IR-Spektroskopie charakterisiert. Eine thermogravimetrische Analyse der Diisopropoxi-Derivate wurde ebenfalls ausgeführt.
  相似文献   

13.
A bimolecular rate constant,kDHO, of (29 ± 9) × 10?12 cm3 molecule?1 s?1 was measured using the relative rate technique for the reaction of the hydroxyl radical (OH) with 3,5‐dimethyl‐1‐hexyn‐3‐ol (DHO, HC?CC(OH)(CH3)CH2CH(CH3)2) at (297 ± 3) K and 1 atm total pressure. To more clearly define DHO's indoor environment degradation mechanism, the products of the DHO + OH reaction were also investigated. The positively identified DHO/OH reaction products were acetone ((CH3)2C?O), 3‐butyne‐2‐one (3B2O, HC?CC(?O)(CH3)), 2‐methyl‐propanal (2MP, H(O?)CCH(CH3)2), 4‐methyl‐2‐pentanone (MIBK, CH3C(?O)CH2CH(CH3)2), ethanedial (GLY, HC(?O)C(?O)H), 2‐oxopropanal (MGLY, CH3C(?O)C(?O)H), and 2,3‐butanedione (23BD, CH3C(?O)C(?O)CH3). The yields of 3B2O and MIBK from the DHO/OH reaction were (8.4 ± 0.3) and (26 ± 2)%, respectively. The use of derivatizing agents O‐(2,3,4,5,6‐pentalfluorobenzyl)hydroxylamine (PFBHA) and N,O‐bis(trimethylsilyl)trifluoroacetamide (BSTFA) clearly indicated that several other reaction products were formed. The elucidation of these other reaction products was facilitated by mass spectrometry of the derivatized reaction products coupled with plausible DHO/OH reaction mechanisms based on previously published volatile organic compound/OH gas‐phase reaction mechanisms. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 534–544, 2004  相似文献   

14.
Bibasic tetradentateSchiff bases having the donor system OH–NX–NX–OH have been shown to form UO2(NO3)2(SBH2) type of derivatives [SBH2 is the molecule of the bibasic tetradentateSchiff bases such as HOC6H4C(R) N(CH2) n NC(R)C6H4OH (where R=H or CH3 andn=2 or 3) and HOC(R)CHC(CH3)N(CH2) n NC(CH3)CH C(R)OH (where R=CH3 or C6H5 andn=2 or 3)]. The 11 stoichiometry of these complexes is shown by elemental analysis and conductometric titrations. The molar conductence values in nitrobenzene indicate the non-electrolytic behaviour and the magnetic susceptibility measurements by the Gouy method show these complexes to be diamagnetic.With 1 Figure  相似文献   

15.
Aqueous solutions of Me2Te(OH)2 and (CH2)4Te(OH)2 readily absorb carbon dioxide giving rise to the formation of the dialkyltelluroxane carbonates (Me2TeOTeMe2CO3)n ( 1 ) and HO(CH2)4TeOTe(CH2)4CO3Te(CH2)4OH·2H2O ( 2 ·2H2O), which were characterised by 13C MAS and 125Te MAS NMR spectroscopy as well as X‐ray crystallography. The spatial arrangement of the tellurium atoms is defined by C2O2 donor sets in the primary coordination sphere and one or two secondary Te···O contacts, which involve coordination of the carbonate moieties. In turn, the different Te–O coordination modes render a lack of symmetry to the carbonate moieties, which show significantly different C–O bond lengths, an important feature when contemplating the C–O bond activation in carbonates. The structural and spectroscopic parameters of 1 and 2 are discussed in comparison with other heavy p‐block element carbonates. In solution, electrolytic dissociation of 1 and 2 takes place.  相似文献   

16.
The differential thermal analyses of n-long chain alcohols CH3(CH3)n-1-OH, designated as Cn-OH (n = 16, 18, 20 and 22), and corresponding alkoxy ethanols CH3(CH2)n-1-OCH2CH2OH, designated as Cn-OC2H4OH, have been carried out at a heating rate of 1°C min?1. The differential e.m.f (ΔV) has been plotted against the temperature of the reference material (°C) and the onset of the peak has been taken as the appearance of the polymorphic phase transition or melting of the compound. Heats of transition (ΔH1) and melting or fusion (ΔHf) were computed from the areas under the respective peaks.  相似文献   

17.
The electron impact induced loss of a phenoxy radical from the molecular ions of α,ω-bis-aryloxy alkanes ΦO(CH2)nOΦ ( 1 n; n = 2–7) and F-p-C6H4(CH2)nOΦ(2n; n = 2?5) is the result of functional group interaction. Labelling data provide conclusive evidence for the O-aryl tetra-hydrofuranium and O-aryl tetrahydropyranium structures of the resulting decomposing species (lifetimes between 10?6 and 10?5 s) in the case of n = 4 and 5, respectively. Evidence is presented for the occurrence of phenyl participation in the loss of ΦOH from the molecular ions of the lower homologues of 1 n and 2n (n = 2, 3).  相似文献   

18.
Reactions of silicon tetraacetate with different types ofSchiff bases have been investigated in anhydrous benzene. Monofunctional bidentate, C6H5CHNXOH and HORCHNC6H5 [whereX=CH2CH2, CH2CH(CH3) or o-C6H4 and R=o-C6H4 or 2,1-C10H6], bifunctional tridentate, o-HOC6H4CHNYOH [whereY=CH2CH2 or CH2CH(CH3)] and bifunctional tetradentateSchiff bases, o-HOC6H4C(CH3)N(CH2) n NC(CH3)C6H4OH-o (wheren=2 or 3) have been shown to yield derivatives of the type, Si(OAc)4– m L m, Si(OAc)4–2 n L n and Si(OAc)2 L (wherem=1,2 or 3;n=1 or 2 and HL, H2 L and H2 L represent the molecules of monofunctional bidentate, bifunctional tridentate and bifunctional tetradentateSchiff bases resp.) and have been found to be monomeric in boiling benzene. Tentative structures based on IR and in a few cases PMR spectra have been indicated for the resulting derivatives.With 2 Figures  相似文献   

19.
The controlled cationic polymerization of cyclopentadiene (CPD) at 20 °C using 1‐(4‐methoxyphenyl)ethanol (1)/B(C6F5)3 initiating system in the presence of fairly large amount of water is reported. The number–average molecular weights of the obtained polymers increased in direct proportion to monomer conversion in agreement with calculated values and were inversely proportional to initiator concentration, while the molecular weight distribution slightly broadened during the polymerization (Mw/Mn ~ 1.15–1.60). 1H NMR analyses confirmed that the polymerization proceeds via reversible activation of the C? OH bond derived from the initiator to generate the growing cationic species, although some loss of hydroxyl functionality happened in the course of the polymerization. It was also shown that the enchainment in cationic polymerization of CPD was affected by the nature of the solvent(s): for instance, polymers with high regioselectivity ([1,4] up to 70%) were obtained in acetonitrile, whereas lower values (around 60%) were found in CH2Cl2/CH3CN mixtures. Aqueous suspension polymerization of CPD using the same initiating system was successfully performed and allowed to synthesize primarily hydroxyl‐terminated oligomers (Fn = 0.8–0.9) with Mn ≤ 1000 g mol?1 and broad MWD (Mw/Mn ~ 2.2). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4734–4747, 2008  相似文献   

20.
Specific magnetic susceptibilities (s) of several newly synthesized chelates of some of the lanthanons [La(III), Pr(III) and Nd(III)] are reported. These derivatives are of the general type,Ln(O-i-C3H7)3–n (C6H5CHNRO) n [where,Ln=La(III), Pr(III) or Nd(III);n=1 or 2 and R=CH2CH2, CH2CHCH3 or C6H4] and have been prepared by the reaction of the alkoxides of the lanthanons withSchiff bases such as benzylidene-2-hydroxyethylamine (C6H5CHNCH2CH2OH), benzylidene-2-hydroxy-n-propylamine (C6H5CHNCH2CHOHCH3) and benzylidene-o-aminophenol (C6H5CHNC6H4OH) in different molar relations in dry benzene.The resulting crystalline derivatives are non-volatile, light to deep yellow or blackish in colour. These tend to polymerize on keeping as shown by their insoluble nature and higher melting points, the polymerisation possibly occurring by the intermolecular coordination through oxygen atoms as reported earlier1.UsingGouy method2, the bis-isopropoxy mono-Schiff base and mono-isopropoxy bis-Schiff base complexes of La(III) have been shown to be diamagnetic, with s values being in the range of –0.32 to –0.45×10–6 and –0.39 to –0.55×10–6 c.g.s. units at 305 K respectively.In the remaining derivatives, Pr(O-i-C3H7)3–n (C6H5CH NRO) n and Nd(O-i-C3H7)3–n (C6H5CHNRO) n (where,n=1 or 2 and R=CH2CH2, CH2CHCH3 or C6H4) the magnetic moment values range between 3.25 to 3.32 and 3.30 to 3.33 B respectively indicating their paramagnetic nature.  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