共查询到20条相似文献,搜索用时 15 毫秒
1.
Gabriel Cendejas Csar Andrs Flores-Sandoval Nelson Huitrn Rafael Herrera Luis S. Zamudio-Rivera Hiram I. Beltrn Flavio Vzquez 《Journal of Molecular Structure》2008,879(1-3):40-52
In this work, the influence of three different initiators (KOH, KOH dissolved in ethanol and the potassium salt of ethylene glycol) on the propylene oxide polymerization was studied by experimental and theoretical methods. A first series of reactions was carried out to establish the adequate thermal conditions for a minimal monomer transfer during the polymerization. The formation of end insaturations (main consequence of the monomer transfer interference) in the poly(propylene oxide) chains was studied by spectroscopic methods. Furthermore, a second series of poly(propylene oxide)s was prepared by using the mentioned initiators, and characterized by size exclusion chromatography. The initiator efficiency to create active centers in every reactive system was determined from the molecular weight and the conversion data obtained. Experimental results were elucidated by using quantum chemical calculations at density functional theory level, involving thermo-chemistry parameters, and the simulation of the infrared, and 13C nuclear magnetic resonance spectra. This method led to studying the addition of up to ten propylene oxide unit, resulting into important energetic tendencies and regioselectivity, being compared to the physicochemical data of products obtained. These correlations meant further understanding of the reaction course and the type of products obtained, depending on the nature of the initiator. 相似文献
2.
Yi‐Jun Huang Guo‐Rong Qi Yu‐Hua Wang 《Journal of polymer science. Part A, Polymer chemistry》2002,40(8):1142-1150
A double metal‐cyanide catalyst based on Zn3[Co(CN)6]2 was prepared. This catalyst is very effective for the ring‐opening polymerization of propylene oxide. Polyether polyols of moderate molecular weight having low unsaturation (<0.015 meq/g) can be prepared under mild conditions. The molecular weight of polymer is entirely controlled by a reacted monomer‐to‐initiator ratio. The polymers prepared with stepwise addition of monomer exhibit a narrower molecular weight distribution as compared with those prepared with one‐step addition of monomer. Various compounds containing active hydrogen, except basic compounds and low‐carbon carboxylic acid, may be used as initiators. The reaction rate increases with increasing catalyst amount and decreases with rising initiator concentration. Polymerization involves a rapid exchange reaction between the active species and the dormant species. It was also proven that, to a certain extent, the chain termination of this catalytic system is reversible or temporary. 13C NMR analysis showed that the polymer has a random distribution of the configurational sequences and head‐to‐tail regiosequence. It is assumed that the polymerization proceeds via a cationic coordination mechanism. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1142–1150, 2002 相似文献
3.
Xianbiao Zeng Yifeng Zhang Zhiquan Shen 《Journal of polymer science. Part A, Polymer chemistry》1997,35(11):2177-2182
Ring opening polymerization of propylene oxide in the presence of a new type of catalytic system composed of chitosan-supported rare earth complex, triisobutyl aluminium, and acetylacetone and its kinetics have been studied for the first time. It has been found that the characteristics of this catalytic system are of high catalytic activity, of higher stereoselectivity, and of a high molecular weight polymer of 2 × 106. Kinetic studies show that the polymerization rate is first order with respect to monomer concentration and catalyst concentration, respectively. The apparent activation energy of the polymerization reaction is 37.1 kJ/mol. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35 : 2177–2182, 1997 相似文献
4.
5.
Victoria Volkis Elena Smolensky Anatoli Lisovskii Moris S. Eisen 《Journal of polymer science. Part A, Polymer chemistry》2005,43(19):4505-4516
The effect of solvents (toluene, dichloromethane, and hexane) was studied on the polymerization of propylene with the octahedral complexes bis(trimethylsilyl)benzamidinate titanium dichloride(a), bis(acetylacetonate) titanium dichloride(b), and bis(diethylamino) titanium di‐2‐(diphenylphosphanylamino)pyridine as catalytic precursors and methylalumoxane as the cocatalyst. For comparison, the polymerization was also performed in plain liquid propylene without the addition of any solvent. The obtained polymers were fractionated by refluxing hexane. The activity of the complexes and the molecular weights and tacticities of the whole polymers and their different fractions were the studied parameters. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 4505–4516, 2005 相似文献
6.
Two new bifunctional salen-cobalt(III) complexes were synthesized, which consist of salicylaldehyde bearing four quaternary ammonium salts and two different diamines. The copolymerization results indicated that decreasing temperature is advantageous for both the complexes. Of both the diamines, the complex 9 with o-diaminobenzene has a higher catalytic effect compared to complex 6 with 1,2-diaminocyclohexane. The catalytic effect of complex 9 is over 3.5 times than that of complex 6 at a temperature of 30°C. The research of PCO2 on the copolymerization revealed that the first-rank pressure was at 2 MPa for the two complexes. The highest turnover number are under conditions of T = 30°C, PCO2 = 2 MPa, and t = 24 hr. Differential scanning calorimeter curves indicated that poly(propylene carbonate) (PPC) by complex 9 has the highest Tg of 54.2°C. DTGA curves showed that there were two thermal degradation peaks, the first is for the ester bond, and the second is for the C–C bond. 相似文献
7.
Sakharov A. M. Il"in V. V. Rusak V. V. Nysenko Z. N. Klimov S. A. 《Russian Chemical Bulletin》2002,51(8):1451-1454
Copolymerization of carbon dioxide with propylene oxide in the presence of zinc adipate was studied. The effects of the temperature, nature of the solvent, and catalyst concentration on the molecular weight, molecular-weight distribution, and yields of the copolymer and propylene carbonate were examined. The structure of the polymer obtained was studied by 13N and 1I NMR spectroscopy. 相似文献
8.
Rainer Koch 《Journal of organometallic chemistry》2006,691(21):4539-4544
A series of six titanium pentafulvene complexes are thoroughly investigated using the B3LYP/6-31G(d) level of theory. Excellent agreement with the available structural data is obtained. Relevant structural parameters indicate that a gradual change of the fulvene ligand coordination to the titanium center. Depending on the nature of the exocyclic fulvene substituents dianionic η5, η1-and olefinic η6-coordination modes are found. This behavior is further supported by NBO and AIM population analyses which predict differences in the bond nature meanly of the contacts between titanium and the exocyclic carbon atom. In an excitation study, several theoretical approaches are evaluated against the available recorded UV-Vis spectra of the six complexes. The “best” approach, time-dependant DFT calculations reproduce the experimental UV data reasonably well, although systematically slightly too small values (abut 50 cm−1) are obtained. The other levels of theory are significantly more erratic. It could further be testified that the absorption maxima correspond to a ligand-to-metal charge transfer from the HOMO to the LUMO+1 of the complexes. 相似文献
9.
10.
Nieto Faza O Silva López C Alvarez R de Lera AR 《Chemistry (Weinheim an der Bergstrasse, Germany)》2004,10(17):4324-4333
At the 6-311G* level of theory, DFT methods predict that the rearrangement of 1,4-dihydroxy-5-methylpentadienyl cation 1 (R = Me) to protonated trans-3-hydroxy-2-methylcyclopent-4-en-1-one 2, an intermediate step in the Piancatelli reaction or rearrangement of furfuryl carbinols to trans-2-alkyl(aryl)-3-hydroxycyclopent-4-en-1-one, is a concerted electrocyclic process. Energetic, magnetic, and stereochemical criteria are consistent with a conrotatory electrocyclic ring closure of the most stable out,out-1 isomer to afford trans-2. Although the out,in-1 isomer is thermodynamically destabilized by 6.84 kcal mol(-1), the activation energy for its cyclization is slightly lower (5.29 kcal mol(-1) versus 5.95 kcal mol(-1)). The cyclization of the isomers of 1 with the C1-hydroxy group inwards showed considerably higher activation energies than their outwards counterparts. in,out-1, although close in energy to out,out-1 (difference of 1.57 kcal mol(-1)) required about 10 kcal mol(-1) more to reach the corresponding transition structure. The value measured for the activation energy of in,in-1 (17.32 kcal mol(-1)) eliminates the alternative conrotatory electrocyclization of this isomer to provide trans-2. Geometric scrambling by isomerization of the terminal C1--C2 bond of 1 is also unlikely to compete with electrocyclization. The possibility to interpret the 1-->2 reaction as a nonpericyclic cationic cyclization was also examined through NBO analysis, and the study of bond lengths and atomic charges. It was found that the 1-->2 concerted rearrangement benefits from charge separation at the cyclization termini, an effect not observed in related concerted electrocyclic processes, such as the classical Nazarov reaction 3-->4 or the cyclization of the isomeric 2-hydroxypentadienyl cation 5. 相似文献
11.
Man ML Lam KC Sit WN Ng SM Zhou Z Lin Z Lau CP 《Chemistry (Weinheim an der Bergstrasse, Germany)》2006,12(4):1004-1015
The heterobimetallic complexes [(eta5-C5H5)Ru(CO)(mu-dppm)Mn(CO)4] and [(eta5-C5Me5)Ru(mu-dppm)(mu-CO)2Mn(CO)3] (dppm = bis-diphenylphosphinomethane) have been prepared by reacting the hydridic complexes [(eta5-C5H5)Ru(dppm)H] and [(eta5-C5Me5)Ru(dppm)H], respectively, with the protonic [HMn(CO)5] complex. The bimetallic complexes can also be synthesized through metathetical reactions between [(eta5-C5R5)Ru(dppm)Cl] (R = H or Me) and Li+[Mn(CO)5]-. Although the complexes fail to catalyze the hydrogenation of CO2 to formic acid, they catalyze the coupling reactions of epoxides with carbon dioxide to yield cyclic carbonates. Two possible reaction pathways for the coupling reactions have been proposed. Both routes begin with heterolytic cleavage of the RuMn bond and coordination of an epoxide molecule to the Lewis acidic ruthenium center. In Route I, the Lewis basic manganese center activates the CO2 by forming the metallocarboxylate anion which then ring-opens the epoxide; subsequent ring-closure gives the cyclic carbonate. In Route II, the nucleophilic manganese center ring-opens the ruthenium-attached epoxide to afford an alkoxide intermediate; CO2 insertion into the RuO bond followed by ring-closure yields the product. Density functional calculations at the B3LYP level of theory were carried out to understand the structural and energetic aspects of the two possible reaction pathways. The results of the calculations indicate that Route II is favored over Route I. 相似文献
12.
Alireza Ariafard 《Journal of organometallic chemistry》2006,691(21):4545-4555
Density functional theory calculations were carried out to investigate structures and stabilities of tropone and troponeiron complexes, (tropone)Fe(CO)3, (tropone)Fe(CO)2(PH3) and (tropone)Fe(PH3)3, and their protonated species. The results show that the oxygen-protonated tropone is more stable than the carbon-protonated tropone. On the contrary, in the troponeiron complexes, the carbon protonated species are more stable than the oxygen protonated species. In the neutral and oxygen-protonated complexes, the tropone and oxygen-protonated tropone ligands are η4-coordinated. In the carbon-protonated complexes, the carbon-protonated tropone ligand is η5-coordinated. The results also show that the metal shift for complexes containing phosphine ligands is more difficult than that for those containing carbonyl ligands. For the neutral methyl-substituted troponeiron complexes, steric effect was found to play a key role in determining the relative stability of the regioisomers. For their protonated species, the electron-donating properties of the methyl substituent(s) were found to be important in determining the relative stability among the different regioiosmers. 相似文献
13.
《先进技术聚合物》2018,29(6):1870-1874
In this study, we have for the first time demonstrated that palladium chloride (PdCl2) is an efficient catalyst for ring‐opening polymerization of cyclohexene oxide in a solvent‐free condition. The polymerization product was in atactic structure, and reaction conditions, such as reaction temperature, time, and catalyst amount, showed effects on polymerization conversion yield, turnover number, and number‐average molecular weight of the resulting poly(cyclohexene oxide). PdCl2 catalysis follows a cationic ring‐opening mechanism. The polymerization result is highly determined by the chemical structure of the monomers. 相似文献
14.
Daisuke Takeuchi Yoshihiko Sakaguchi Kohtaro Osakada 《Journal of polymer science. Part A, Polymer chemistry》2002,40(24):4530-4537
Co2(CO)8 catalyzes the ring‐opening copolymerization of propylene oxide with CO to afford the polyester in the presence of various amine cocatalysts. The 1H and 13C{1H} NMR spectra of the polyester, obtained by the Co2(CO)8–3‐hydroxypyridine catalyst, show the following structure ? [CH2? CH(CH3)? O? CO]n? . The Co2(CO)8–phenol catalyst gives the polyester, which contains the partial structural unit formed through the ring‐opening copolymerization of tetrahydrofuran with CO. The bidentate amines, such as bipyridine and N,N,N′,N′‐tetramethylethylenediamine, enhance the Co complex‐catalyzed copolymerization, which produces the polyester with a regulated structure. Acylcobalt complexes, (RCO)Co(CO)n (R = Me or CH2Ph), prepared in situ, do not catalyze the copolymerization even in the presence of pyridine. This suggests that the chain growth involves the intermolecular nucleophilic addition of the OH group of the intermediate complex to the acyl–cobalt bond, forming an ester bond rather than the insertion of propylene oxide into the acyl–cobalt bond. Co2(CO)8? Ru3(CO)12 mixtures also bring about the copolymerization of propylene oxide with CO. The molar ratio of Ru to Co affects the yield, molecular weight, and structure of the produced copolymer. The catalysis is ascribed to the Ru? Co mixed‐metal cluster formed in the reaction mixture. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4530–4537, 2002 相似文献
15.
L. P. Kosmacheva R. F. Ambartsumova E. G. Mil'grom 《Chemistry of Heterocyclic Compounds》2000,36(8):993-995
In reaction of 2-aminobenzoxazole with propylene oxide in protic solvents, opening of the heterocycle ring occurs along with alkylation. In the case of 2-aminobenzimidazole, only alkylation occurs under similar conditions. We have studied the compositions of the reaction mixtures and the dynamics of formation of the reaction products by mass spectrometry.Institute of Plant Substances Chemistry, Academy of Sciences of the Uzbekistan Republic, Tashkent 700170. Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 8, pp. 1132–1134, August, 2000. 相似文献
16.
Mercero JM Formoso E Matxain JM Eriksson LA Ugalde JM 《Chemistry (Weinheim an der Bergstrasse, Germany)》2006,12(17):4495-4502
We report on novel sandwichlike structures [Al(4)MAl(4)](q-) (q=0-2 and M=Ti, V, Cr, Zr, Nb, Mo, Hf, Ta and W) based on the recently synthesized all-metal aromatic Al(4)(2-) square ring. The sandwichlike structures have two aromatic tetraaluminum square rings which trap a transition-metal cation from either the first, second, or third row. The stability of the anionic sandwichlike complexes towards electron detachment is discussed, and addition of alkali cations is found to stabilize the 2- charged complexes, preventing spontaneous electron detachment. Once the sandwichlike complexes are formed, the Al(4)(2-) square properties remain nearly unchanged; this fact strongly supports the hypothesis that in these complexes the Al(4)(2-) square rings remain aromatic. 相似文献
17.
Il Kim Seung Hoon Byun Chang‐Sik Ha 《Journal of polymer science. Part A, Polymer chemistry》2005,43(19):4393-4404
Polymerizations of propylene oxide were carried out with double metal cyanide (DMC) catalysts based on Zn3[Co(CN)6]2. Through the control of the type and amount of ZnX2 (X = F, Cl, Br, or I) during the preparation of the catalyst, the catalytic activity, induction period, and unsaturation level in the polyether polyols could be tuned. The DMC catalysts were characterized by X‐ray photoelectron spectroscopy, infrared spectroscopy, and X‐ray powder diffraction. In general, ZnBr2 was the most effective zinc halide with respect to the properties of the resulting polymers as well as the activity and induction period. The average rates of polymerizations of DMC catalysts prepared with ZnCl2, ZnBr2, and ZnI2 were 889, 1667, and 784 g of polyoxypropylene/g of catalyst h, respectively, with induction periods of about 53, 5, and 60 min, respectively, at 115 °C. The DMC catalysts produced polyoxypropylenes with an ultralow unsaturation level (0.0025–0.0057 mequiv/g) and a narrow molecular weight distribution (1.07–1.42) without high‐molecular‐weight tails; this resulted in a low viscosity (962–3950 cP). According to the results collected from catalyst characterizations and polymerizations, the active sites of DMC‐catalyzed polymerization had mainly coordinative characters. The presence of free anions accelerated the ring‐opening procedure and thus enhanced the propagation rate and shortened the induction period. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 4393–4404, 2005 相似文献
18.
Bo Li Guang‐Peng Wu Wei‐Min Ren Yi‐Ming Wang Dun‐Yan Rao Xiao‐Bing Lu 《Journal of polymer science. Part A, Polymer chemistry》2008,46(18):6102-6113
Chiral chromium complexes of tetradentate N,N′‐disubstituted bis(aminophenoxide) (designated as Salan, a saturated version of Schiff‐base Salen ligand) in conjunction with an ionic quaternary ammonium salt can efficiently catalyze the copolymerization of CO2 with racemic propylene oxide (rac‐PO) at mild conditions to selectively afford completely alternating poly(propylene carbonate) (PPC) with ~ 95% head‐to‐tail linkages and moderate enantioselectivity. These new catalyst systems predominantly exceed the previously much‐studied SalenCr(III) systems in catalytic activity, polymer enantioselectivity, and stereochemistry control. The chiral diamine backbone, sterically hindered substitute groups on the aromatic rings, and the presence of sp3‐hydridized amino donors and its N,N′‐disubstituted groups in chiral SalanCr(III) complexes all play significant roles in controlling polymer stereochemistry and enantioselectivity. Furthermore, a relationship between polycarbonate enantioselectivity and its head‐to‐tail linkages in relation to regioselective ring‐opening of the epoxide was also discussed on the basis of stereochemical studies of PPCs derived from the copolymerization of CO2 with chiral PO at various conditions. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 6102–6113, 2008 相似文献
19.
Suraj Bhunora Jane Mugo Archana Bhaw‐Luximon Selwyn Mapolie Juanita Van Wyk James Darkwa Ebbe Nordlander 《应用有机金属化学》2011,25(2):133-145
A range of monomeric tetra‐coordinate copper (II) and zinc (II) complexes based on N,O‐bidentate salicylaldimine Schiff base ligands has been synthesized and characterized using various spectroscopic techniques. These complexes were then evaluated as initiators in ring‐opening polymerization of lactides at both 70 °C and 110 °C. The effect of structural changes in the complexes on the ability of these compounds to initiate lactide polymerization as well as the impact on the chemical and physical characteristics of the polymers obtained indicate that the coordination geometry of the metal complex, M? O bond length and substituents on the Schiff base ligand all play a role in the catalyst activity. Electronic factors were dominant in the case of the copper complexes while steric factors prevailed in the case of Zn initiators. Both the Zn and Cu complexes exhibit characteristics of living ring opening polymerization. Copyright © 2010 John Wiley & Sons, Ltd. 相似文献
20.
Zexing Cao Xi Jin Zhaohui Zhou Qianer Zhang 《International journal of quantum chemistry》2006,106(9):2161-2168
Protonation of the metal‐bound oxy‐bidentate ligand in the model complexes of [(HS)3(NH3)M(OCH2COO)]q (M = Mo, Fe, V, Co; q = ?2, ?1) in the gas phase and in solutions of water and acetonitrile has been explored by the density functional approach. Calculations show that protonation of the carboxyl oxygen can open the α‐hydroxycarboxylate chelate ring ligated to a transition‐metal center under specific oxidation and spin states. The feasibility of the chelate ring opening by protonation depends on the electronic nature of the metal site in tune with conversion of a six‐coordinate with a five‐coordinate metal atom. Such selective dissociation of the metal‐bound chelate ligand manipulates the availability of an empty site at the metal center and significantly affects reactivity of the metal‐mediated chemical processes. Protonation changes the stability of species with different spin multiplicities and impels spin transition at the metal center in dissociation of the oxy‐bidentate ligand. Solvent environments of water and acetonitrile play an important role in stabilizing the negatively charged species. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006 相似文献