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 500 毫秒
1.
The present study details the highly effective and controlled ring-opening polymerization (ROP) of α-chloro-ε-caprolactone ( 1 , αClεCL), a cyclic ester that has been little explored thus far in ROP catalysis, using Zn- and Al-based catalysts [Zn(C6F5)2(toluene)] ( 4 ), [N,N′-bis(3,5-di-tert-butylsalicylidene)1,3-diaminopropanato]aluminium(III)benzyloxide ( 5 ) and [N,N′-bis(3,5-di-tert-butylsalicylidene)1,3-diamino-2,2′-dimethylpropanato]aluminium(III)benzyloxide] ( 6 ). Chain-length-controlled PαClεCL material is produced under solution ROP conditions, as deduced from GPC, NMR, MALDI-TOF, and kinetic data. In contrast, the ROP of 1 is ill-defined under bulk ROP conditions due to partial thermal degradation of the polymer chain (presumably through C–Cl cleavage), reflecting the limited stability of PαClεCL. The Al Catalysts 5 and 6 are highly active ROP catalysts of αClεCL at room temperature (TOF up to 2,400 hr−1) to afford well-defined P(αClεCL). In the case of Catalyst 6 , carrying out the ROP of αClεCL under immortal conditions (with BnOH as chain transfer agent) is clearly beneficial to ROP activity and control, with no apparent side-reaction of chloro-functionalized PCL chains as the ROP proceeds. The controlled character of these ROPs was further exploited for the production of chain-length-controlled PLLA-b-PαClεCL diblocks through sequential ROP of l -lactide and αClεCL, affording copolymers with improved thermal and biodegradable properties.  相似文献   

2.
This study compared ε-caprolactone (CL), L-lactide (LA), and ε-decalactone (DL) polymerizations, where aluminum complexes bridged by two pyrazole ligands was used as catalysts. The reactivites of these Al complexes between CL, LA, and DL polymerization were different that L Bu 2 Al 2 Me 4 , with the distort boat form, exhibits the greatest catalytic activity for LA and DL polymerization at 60°C but the lowest catalytic activity for CL polymerization at room temperature. This may be because dinuclear L Bu 2 Al 2 Me 4 cannot react with BnOH to form aluminum benzyl oxide at room temperature, making it unable to reduce catalytic activity. Because these aluminum complexes had different reactivities for LA and CL polymerizations, the selectivity of polylactide-gradual-poly-ε-caprolactones (PLA(10–80%)-gradual-PCL(59–79%)) (PLA-g-PCLs) was observed.  相似文献   

3.
An effective approach for fast synthesis of well-defined star-shaped poly(2-methyl-N-tosylaziridine)s was developed by one-pot tandem ring-opening polymerization (ROP) of N-sulfonyl aziridines with trimethylsilyl azide (TMSN3) and “click” reaction with alkynes. Azido terminated polyaziridines (α-N3-PAzs) could be achieved via ROP of N-sulfonyl aziridines with TMSN3 in the presence of organic superbases. The catalytic efficiency of organobases, including 1,4-diazabicyclo[2.2.2]octane (DABCO), 1,8-diazabicyclo[5.4.0]-7-undecene (DBU), 7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene (MTBD), 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD), and N,N,N′,N″,N″-pentamethyldiethylenetriamine (PMDETA), was evaluated, and all of them except TBD afforded “living”/controlled ROP of 2-methyl-N-tosylaziridines (TsMAz). Star-shaped polyaziridines were then fastly synthesized by the one-pot tandem strategy. During the reaction process, PMDETA catalyzed ROP first, then was triggered to be a ligand by adding CuBr for “click” reaction. Well-defined 3- and 4-arm star P(TsMAz)s were successfully prepared, and subsequently desulfonylated to give star-shaped polypropylenimines (PPIs). PPI stars exhibited intrinsic photoluminescence properties from the polyamine arms.  相似文献   

4.
《Mendeleev Communications》2021,31(6):850-852
Polydentate SNO-coordinating proligands were obtained by addition of thioacetic acid to conjugated alkene, or ring-opening of thiiranes with organolithium compounds as the key stages. These SNO- and known SNS- and SOS- coordinating proligands were used for the synthesis of tetrylenes by the reaction with Lappert’s tetrylenes E[N(SiMe3)2]2 (E = Ge, Sn), giving polymeric, monomeric or dimeric species depending on the type of the ligand. The structure of stannylenes in solution was analysed by 119Sn NMR spectroscopy.  相似文献   

5.
Two novel aluminum metal complexes ( 2 and 3 ) bearing salen ligands were in situ prepared from trimethyl aluminum (AlMe3), methanol, and (R,R)‐N,N′‐bis(salicylidene)‐1,2‐diaminocyclohexane with original synthetic strategies, and a preliminarily resoluted (R,R)‐1,2‐diaminocyclohexane was applied as a synthetic precursor. By means of Fourier transform infrared spectrometry, NMR spectrometry, mass spectrometry, and single‐crystal X‐ray diffractometry, 2 and 3 were revealed to be distinct molecular structures with corresponding yields of 85 and 10%, respectively. Further studies via 27Al NMR techniques and single‐crystal X‐ray diffraction indicated that dimeric metal complex 3 appeared in the six‐coordinated state, whereas there was a dynamic equilibrium transition between the five‐ and six‐coordinated states for metal complex 2 in a CDCl3 solution. The more stable dimeric metal complex ( 3 ) exhibited two inequivalent aluminum metal centers coordinated to nitrogen atoms attributed to two different salen ligands, and this was different from the reported salen aluminum complex structures. Furthermore, 2 and 3 were employed as candidate catalysts for the ring‐opening polymerization (ROP) of some important biodegradable aliphatic polyesters and polycarbonates, including poly(?‐caprolactone) (PCL), poly(δ‐valerolactone), poly(trimethylene carbonate), and poly(2,2‐dimethyl trimethylene carbonate). The synthetic results indicated that both metal complexes efficiently catalyzed ROP at 100 °C in an anisole solution, and 3 showed much better controlled characteristics of ROP than 2 . Very narrow molecular weight distributions close to 1.21 for PCL were detected with 3 as the ROP catalyst. In addition, a catalytic mechanism study confirmed that ROP catalyzed by these metal complexes was in good agreement with the commonly accepted coordination polymerization reported for aluminum triiso [Al(OiPr)3] and stannous octanoate [Sn(Oct)2]. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 373–384, 2005  相似文献   

6.
Three aluminum complexes supported by salen ligands derived from cis-1,2-cyclohexanediamine and salicylaldehyde derivatives were synthesized. They were characterized by 1H, 13C NMR spectra, and elemental analysis. X-ray diffraction analysis revealed that aluminum was in distorted square pyramidal geometry in 2. These complexes were employed as catalysts for the ring-opening polymerization (ROP) of L-lactide and rac-lactide. Complex 2 showed the highest activity among these complexes with isopropanol for the ROP of L-lactide and 3 showed the highest stereoselectivity for the ROP of rac-lactide attaining partially isotactic polylactide with a Pm of 0.75. The kinetic data of the polymerization utilizing 3 as catalyst showed that the polymeric rate was first order to the monomer and catalyst.  相似文献   

7.
A series of novel diaroylhydrazone aluminum complexes have been synthesized and well-defined structurally, and their catalytic performance in the polymerization of ε-caprolactone and lactides have also been evaluated. Complexes [(L1–4)2AlMe] ( 1 – 4 ) {[L1 = (3,5-tBu2–2-OMe-C6H2)CH=NNCOC6H5], [L2 = (3,5-tBu2–2-OMe-C6H2)CH=NNCO(C6H4–4-OCH3)], [L3 = (3,5-tBu2–2-OMe-C6H2)CH=NNCO(C6H4–4-Br)] and [L4 = (2-OMe-C6H4)CH=NNCO(C6H4–4-tBu)]} were prepared through treatment of AlMe3 with the corresponding proligands L1–4H in molar ratios of 1: 1 or 1: 2. Chemical structures of all the complexes were well-defined by elemental analysis, NMR spectra as well as single-crystal X-ray study. Complexes [(L1–4)2AlMe] ( 1 – 4 ) in this work represent the first examples of aluminum complexes of aroylhydrazone ligands with crystallographic characterization. Specifically, they are all in monomeric form with a penta-coordinated aluminum center, including two approximately co-planar five-membered metallacycles with aluminum. Introduced bulky tert-butyl substituents in aroylhydrazone ligands could affect the geometry around the central metal which is a distorted square-based pyramid in complexes 1 – 3 while being a trigonal bipyramidal in complex 4 , thus affecting their catalytic behaviors. The complexes can successfully catalyze the ring-opening polymerization of ε-caprolactone and L-lactide under mild conditions without any activator. In addition, complexes 1 – 4 could also polymerize rac-lactide, affording atactic polylactides with high conversions and good controllability in relatively short reaction time.  相似文献   

8.
采用溶胶凝胶法合成了一系列有序性好且酸性较强的介孔硅铝酸盐材料。利用X射线粉末衍射(XRD)、透射电镜(TEM)、27Al核磁共振(27Al NMR)、氨气程序升温脱附(HN3-TPD)及吡啶吸附红外光谱(Py-FT-IR)对制备的介孔硅铝酸盐材料的结构和性能进行表征,并考察了材料在苯甲醚和苯甲醇的傅克烷基化反应中的催化活性。实验结果表明:合成过程中,表面活性剂的用量、硅铝物质的量之比会影响材料结构的有序性,醋酸用量对材料结构有序性影响很小;进一步研究结果表明,nSi/nAl比会影响材料的酸催化活性,当nSi/nAl=10时材料的酸催化活性最高。氨气程序升温脱附和吡啶吸附红外光谱表明nSi/nAl=10的材料含有最多的B酸酸量。  相似文献   

9.
《Mendeleev Communications》2021,31(6):847-849
New SNS-ligands were obtained by consequent ring-opening of substituted thiiranes by lithiated 2,6-lutidine, mono- functionalized SN-ligands having been isolated as the intermediate compounds. The molecular structure of ligand 2,6-Py(CH2CH2CBn2SH)2 was elucidated by XRD analysis.The reaction of AlMe3 with SNS-ligands afforded monomeric methyl aluminum complexes which have been tested in ring- opening polymerization of ε caprolactone in bulk.  相似文献   

10.
In this contribution, we explored the catalytic ring‐opening polymerization (ROP) of (macro)lactones using salen–aluminum complexes bearing cyclic β‐ketoiminato ligand. The effects of bridge moiety and ring size in the benzocyclane skeleton on the catalytic activity of these complexes were thoroughly investigated. Complex 5 with 2,2‐dimethylpropylene bridge and five‐membered cyclane ring can efficiently catalyze the ROP of ω‐pentadecalactone (ω‐PDL), showing higher catalytic activity (turnover frequency [TOF] up to 309.2 h?1) than the typical Al‐salen analogs bearing salicylaldiminato ligand (TOF = 227.2 h?1). Thus, polyethylene‐like polyester with high‐molecular weight (up to 164.5 kg/mol) could be easily prepared under optimal conditions. In addition, complex 5 can also catalyze the ROP of lactide (LA) and ε‐caprolactone (ε‐CL) with extremely high activity (TOF is high up to 147.6 h?1 and 4752 h?1, respectively). Here, we demonstrated a rare mono‐nuclear salen‐Al complex that can prompt the ROP of (macro)lactones with unprecedentedly high efficiency. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019 , 57, 973–981  相似文献   

11.
The disproportionation of AlCl(THF)n (THF is tetrahydrofuran) in the presence of lithium amidinate species gives aluminium(III) amidinate complexes with partial or full chloride substitution. Three aluminium amidinate complexes formed during the reaction between aluminium monochloride and lithium amidinates are presented. The homoleptic complex tris(N,N′‐diisopropylbenzimidamido)aluminium(III), [Al(C13H19N2)3] or Al{PhC[N(i‐Pr)]2}3, (I), crystallizes from the same solution as the heteroleptic complex chloridobis(N,N′‐diisopropylbenzimidamido)aluminium(III), [Al(C13H19N2)2Cl] or Al{PhC[N(i‐Pr)]2}2Cl, (II). Both have two crystallographically independent molecules per asymmetric unit (Z′ = 2) and (I) shows disorder in four of its N(i‐Pr) groups. Changing the ligand substituent to the bulkier cyclohexyl allows the isolation of the partial THF solvate chloridobis(N,N′‐dicyclohexylbenzimidamido)aluminium(III) tetrahydrofuran 0.675‐solvate, [Al(C19H27N2)2Cl]·0.675C4H8O or Al[PhC(NCy)2]2Cl·0.675THF, (III). Despite having a twofold rotation axis running through its Al and Cl atoms, (III) has a similar molecular structure to that of (II).  相似文献   

12.
Recently O‐carboxyanhydrides (OCAs) have emerged as a class of viable monomers which can undergo ring‐opening polymerization (ROP) to prepare poly(α‐hydroxyalkanoic acid) with functional groups that are typically difficult to achieve by ROP of lactones. Organocatalysts for the ROP of OCAs, such as dimethylaminopyridine (DMAP), may induce undesired epimerization of the α‐carbon atom in polyesters resulting in the loss of isotacticity. Herein, we report the use of (BDI‐IE)Zn(OCH(CH3)COOCH3) ((BDI)Zn‐1, (BDI‐IE)=2‐((2,6‐diethylphenyl)amino)‐4‐((2,6‐diisopropylphenyl)imino)‐2‐pentene), for the controlled ROP of various OCAs without epimerization. Both homopolymers and block copolymers with controlled molecular weights, narrow molecular weight distributions, and isotactic backbones can be readily synthesized. (BDI)Zn‐1 also enables controlled copolymerization of OCAs and lactide, facilitating the synthesis of block copolymers potentially useful for various biomedical applications. Preliminary mechanistic studies suggest that the monomer/dimer equilibrium of the zinc catalyst influences the ROP of OCAs, with the monomeric (BDI)Zn‐1 possessing superior catalytic activity for the initiation of ROP in comparison to the dimeric (BDI)Zn complex.  相似文献   

13.
Abstract

The polymerization of cyclohexene oxide (CHO) by various organoaluminum compounds such as R3Al (R = Me, Et, i-Bu), Et2AlCl, and EtAlCl2 is reported. Ethyl-substituted aluminum compounds were found to be very effective for this polymerization. As more chlorine atom is substituted on the aluminum atom, the polymer yield was increased though the molecular weight was decreased. The polymer yields at varying monomer-to-catalyst mole ratios (M/C) were similar (80–93%). The temperature and solvent effect for the present polymerization were also studied. The present poly(CHO) was a less stereoregular (atactic) isomer, regardless of catalysts and polymerization conditions. The resulting poly(CHO) was a white powder ad was soluble in aromatic and halogenated hydrocarbon solvents such as benzene, chlorobenzene, CCl4, chloroform, etc.  相似文献   

14.
Two hyperbranched bisphosphinoamine (PNP) ligands and chromium complexes were synthesized in good yield with 1.0 generation (1.0 G) hyperbranched macromolecules, chlorodiphenylphosphine (Ph2PCl) and CrCl3(THF)3 as raw materials. The hyperbranched PNP ligands and chromium complexes were characterized by FT-IR, 1H NMR, 31P NMR, UV and ESI-MS. Comparing with the chromium complexes, the hyperbranched PNP ligands, in combination with Cr(III), and activation by methylaluminoxane (MAO) in situ generated species with better catalytic performance for ethylene oligomerization. The effect of solvent, chromium source, ligand/Cr molar ratio, reaction temperature, Al/Cr molar ratio and reaction pressure on the catalytic activity and product selectivity were studied. The results showed that with increase of ligand/Cr molar ratio, reaction temperature and Al/Cr molar ratio, the catalytic activity increased at first and then decreased. However, the catalytic activity continuously increased with increase of reaction pressure. Under the optimized conditions, the catalytic system of hyperbranched PNP/Cr(III)/MAO led to catalytic activity of 2.68 × 105 g/(mol Cr·h) and 37.71% selectivity for C6 and C8.  相似文献   

15.
The nature behind the promotional effect of phosphorus on the catalytic performance and hydrothermal stability of zeolite H‐ZSM‐5 has been studied using a combination of 27Al and 31P MAS NMR spectroscopy, soft X‐ray absorption tomography and n‐hexane catalytic cracking, complemented with NH3 temperature‐programmed desorption and N2 physisorption. Phosphated H‐ZSM‐5 retains more acid sites and catalytic cracking activity after steam treatment than its non‐phosphated counterpart, while the selectivity towards propylene is improved. It was established that the stabilization effect is twofold. First, the local framework silico‐aluminophosphate (SAPO) interfaces, which form after phosphatation, are not affected by steam and hold aluminum atoms fixed in the zeolite lattice, preserving the pore structure of zeolite H‐ZSM‐5. Second, the four‐coordinate framework aluminum can be forced into a reversible sixfold coordination by phosphate. These species remain stationary in the framework under hydrothermal conditions as well. Removal of physically coordinated phosphate after steam‐treatment leads to an increase in the number of strong acid sites and increased catalytic activity. We propose that the improved selectivity towards propylene during catalytic cracking can be attributed to local SAPO interfaces located at channel intersections, where they act as impediments in the formation of bulky carbenium ions and therefore suppress the bimolecular cracking mechanism.  相似文献   

16.
The preparation of tetradentate amine-bis(phenol) proligands with dichloro and difluoro substituted phenol groups and their reaction with FeX3 (X = Cl or Br) is described. The compounds, 2-pyridylamino-N,N-bis(2-methylene-4,6-dichlorophenol), H2[L1]; 2-pyridylamino-N,N-bis(2-methylene-4,6-difluorophenol), H2[L2]; dimethylaminoethylamino-N,N-bis(2-methylene-4,6-dichlorophenol), H2[L3]; 2-tetrahydrofurfuryl-N,N-bis(2-methylene-4,6-dichlorophenol), H2[L4]; and methoxyethylamino-N,N-bis(2-methylene-4,6-dichlorophenol), H2[L5] were prepared in aqueous medium and obtained as white powders in good to excellent yield. Ten new iron(III) halide complexes supported by these tetradentate ligands are reported. Representative single crystal X-ray diffraction structures were obtained for H2[L1] and a water adduct of the iron(III) complex, aquachloro{2-pyridylamino-N,N-bis(2-methylene-4,6-dichlorophenolato)}iron(III), 2·H2O. The structure of the proligand H2[L1] shows intramolecular hydrogen bonding. In the solid-state structure, the iron complex exhibits intermolecular hydrogen bonding between the water ligand and the phenolate oxygen of a neighbouring complex. The anhydrous complexes were studied for catalytic activity towards C-C cross-coupling of Grignard reagent nucleophiles with alkyl halide electrophiles.  相似文献   

17.
Additions of beryllium–halide bonds in the simple beryllium dihalide adducts, [BeX2(tmeda)] (X=Br or I, tmeda=N,N,N′,N′‐tetramethylethylenediamine), across the metal center of a neutral aluminum(I) heterocycle, [:Al(DipNacnac)] (DipNacnac=[(DipNCMe)2CH]?, Dip=2,6‐diisopropylphenyl), have yielded the first examples of compounds with beryllium–aluminum bonds, [(DipNacnac)(X)Al‐Be(X)(tmeda)]. For sake of comparison, isostructural Mg–Al and Zn–Al analogues of these complexes, viz. [(DipNacnac)(X)Al‐M(X)(tmeda)] (M=Mg or Zn, X=I or Br) have been prepared and structurally characterized. DFT calculations reveal all compounds to have high s‐character metal–metal bonds, the polarity of which is consistent with the electronegativities of the metals involved. Preliminary reactivity studies of [(DipNacnac)(Br)Al‐Be(Br)(tmeda)] are reported.  相似文献   

18.
The ring-opening copolymerization (ROCOP) of functional 4-alkoxymethylene-β-propiolactones (BPLORs) by yttrium-bisphenolate complexes was investigated. The ROCOP of equimolar mixtures of BPLORs of opposite configurations, namely (R)-BPLOR1/(S)-BPLOR2 [R1, R2=OMe, OAllyl, OCH2Ph (=OBn), OSiMe2tBu (=OTBDMS)], by the syndioselective Y{ONOOcum}/iPrOH catalyst/initiator system affords P(HBOR1-alt-HBOR2) copolymers with high alternation degrees (altern.=89–94 %), as determined by comprehensive kinetic, 13C{1H} NMR spectroscopy, MALDI-ToF MS and ESI MS/MS fragmentation studies. The ROCOP of the (R)-BPLOMe/(S)-BPLOTBDMS combination, featuring a large difference in the substituents’ bulkiness, constitutes the only observed exception to this trend (altern.=64 %). On the other hand, the isoselectivity of the Y{ONNOCl}/iPrOH catalyst/initiator system has been exploited to generate, in a one-pot/one-step procedure, original mixtures of isotactic poly(hydroxyalkanoate)s (PHAs). This system efficiently transforms equimolar mixtures of (R)-BPLOAll/(S)-BPLOMe into a 1:1 mixture of the corresponding isotactic iso-(R)-PHBOAll and iso-(S)-PHBOMe homopolymers; almost no copolymerization defects are observed. This new approach has been extended successfully to the ROCOP of equimolar mixtures of racemic monomers, rac-BPLOAll/rac-BPLOMe.  相似文献   

19.
A new series of penta-coordinated Co(III)-, Fe(III)-, and Cr(III)- complexes (610) bearing N-salicylidineisopropylaniline and sodium N-(4-sulfonitsalicylidineisopropyl-aniline) ligands has been synthesized and utilized, after activation with methyl aluminoxane, as catalysts for the polymerization of tert-butylacrylate (t-BA). High molar mass P(t-BA) polymers with very low molecular weight distributions were produced (M w/M n = 1.06–1.09). Cobalt- and chromium-based precatalysts showed higher activity towards the polymerization reaction than those of the iron complexes. The presence of sulfonated groups on the para position of the aryl group in the backbone of the ligand decreases the catalytic activity of the complexes. The ortho alkyl substituents on the aryl groups of the ligand have a favorable influence on the polymerization activity compared to the alkyl-free analogue (11).  相似文献   

20.
The effective catalytic activity of organoaluminum compounds for the monohydroboration of carbodiimides has been demonstrated. Two aluminum complexes, 2 and 3 , were synthesized and characterized. The efficient catalytic performances of four aluminum hydride complexes L1AlH2 (L1=HC(CMeNAr)2, Ar=2,6-Et2C6H3; 1 ), L2AlH2(NMe3) (L2=o-C6H4F(CH=N-Ar), Ar=2,6-Et2C6H3; 2 ), L3AlH (L3=2,6-bis(1-methylethyl)-N-(2-pyridinylmethylene)phenylamine; 3 ), and L4AlH(NMe3) (L4=o-C6H4(N-Dipp)(CH=N-Dipp), Dipp=2,6-iPr2C6H3; 4 ), and an aluminum alkyl complex L1AlMe2 ( 5 ) were used for the monohydroboration of carbodiimides investigated under solvent-free and mild conditions. Compounds 1 – 3 and 5 can produce monohydroborated N-borylformamidine, whereas 4 can afford the C-borylformamidine product. A suggested mechanism of this reaction was explored, and the aluminum formamidinate compound 6 was characterized by single-crystal X-ray, also a stoichiometric reaction was investigated.  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