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 599 毫秒
1.
Anion metathesis of imidazol(in)ium chlorides with KHCO(3) afforded an easy one step access to air stable imidazol(in)ium hydrogen carbonates, denoted as [NHC(H)][HCO(3)]. In solution, these compounds were found to be in equilibrium with their corresponding imidazol(in)ium carboxylates, referred to as N-heterocyclic carbene (NHC)-CO(2) adducts. The [NHC(H)][HCO(3)] salts were next shown to behave as masked NHCs, allowing for the NHC moiety to be readily transferred to both organic and organometallic substrates, without the need for dry and oxygen-free conditions. In addition, such [NHC(H)][HCO(3)] precursors were successfully investigated as precatalysts in two selected organocatalyzed reactions of molecular chemistry and polymer synthesis, namely, the benzoin condensation reaction and the ring-opening polymerization of d,l-lactide, respectively. The generation of NHCs from [NHC(H)][HCO(3)] precursors occurred via the formal loss of H(2)CO(3)via a concerted low energy pathway, as substantiated by Density Functional Theory (DFT) calculations.  相似文献   

2.
Copper and silver N‐heterocyclic carbene (NHC) complexes were prepared through a simple, base‐free protocol involving the decomposition of corresponding imidazol(in)ium‐2‐carboxylates under thermolytic conditions and a subsequent reaction of the in situ generated carbenes with copper(I) or silver(I) chloride, respectively. The desired NHC metal complexes were isolated with good yields after simple crystallization.  相似文献   

3.
A facile synthetic route to N-polyfluoro(trimethylsilyl)ethyl azole derivatives was developed starting from N-bromo(chloro)polyfluoroethyl-substituted azoles. The silanes thus obtained were reacted with various electrophiles in the presence of the fluoride ion to yield the corresponding fluorinated carbinols, ketones, carboxylic acids, and methyl dithiocarboxylates as well as N-pentafluoroethylbenzimidazole.  相似文献   

4.
Five imidazol(in)ium-2-carboxylates bearing cyclohexyl, mesityl, or 2,6-diisopropylphenyl substituents on their nitrogen atoms were prepared from the corresponding N-heterocyclic carbenes (NHCs) by reaction with carbon dioxide. They were characterized by IR and NMR spectroscopies, and by TGA. Their ability to act as NHC precursors for in situ catalytic applications was probed in ruthenium-promoted olefin metathesis and cyclopropanation reactions. When visible light induced ring-opening metathesis polymerization of cyclooctene or cyclopropanation of styrene with ethyl diazoacetate were carried out at 60 °C in the presence of [RuCl2(p-cymene)]2, the NHC · CO2 adducts and their NHC · HX counterparts (X = Cl, BF4) displayed similar activities. When metathesis polymerizations were performed at room temperature, the carboxylates proved far superior to the corresponding imidazol(in)ium acid salts. They displayed the same level of activity as the preformed RuCl2(p-cymene)(IMes) complex, whereas the combination of NHC · HX and KO-t-Bu were almost totally inactive. Results obtained for cyclopropanation reactions at room temperature did not show such a large discrepancy of behavior between the two types of adducts.  相似文献   

5.
N‐Heterocyclic carbene (NHC) complexes bromo(1,3‐dibenzyl‐1,3‐dihydro‐2H‐imidazol‐2‐ylidene)silver(I) ( 2a ), bromo[1‐(4‐cyanobenzyl)‐3‐methyl‐1,3‐dihydro‐2H‐imidazol‐2‐ylidene]silver(I) ( 2b ), and bromo[1‐(4‐cyanobenzyl)‐3‐methyl‐1,3‐dihydro‐2H‐benzimidazol‐2‐ylidene]silver(I) ( 2c ) were prepared by the reaction of 1,3‐dibenzyl‐1H‐imidazol‐3‐ium bromide ( 1a ), 3‐(4‐cyanobenzyl)‐1‐methyl‐1H‐imidazol‐3‐ium bromide ( 1b ), and 3‐(4‐cyanobenzyl)‐1‐methyl‐1H‐benzimidazol‐3‐ium bromide ( 1c ), respectively, with silver(I) oxide. NHC Complexes chloro(1,3‐dibenzyl‐1,3‐dihydro‐2H‐imidazol‐2‐ylidene)gold(I) ( 3a ), chloro[1‐(4‐cyanobenzyl)‐3‐methyl‐1,3‐dihydro‐2H‐imidazol‐2‐ylidene]gold(I) ( 3b ), and chloro[1‐(4‐cyanobenzyl)‐3‐methyl‐1,3‐dihydro‐2H‐benzimidazol‐2‐ylidene]gold(I) ( 3c ) were prepared via transmetallation of corresponding (bromo)(NHC)silver(I) complexes with chloro(dimethylsulfido)gold(I). The complex 3a was characterized in two polymorphic forms by single‐crystal X‐ray diffraction showing two rotamers in the solid state. The cytotoxicities of all three bromo(NHC)silver(I) complexes and three (chloro)(NHC)gold(I) complexes were investigated through 3‐(4,5‐dimethylthiazol‐2‐yl)‐2,5‐diphenyl‐2H‐tetrazolium bormide (MTT)‐based preliminary in vitro testing on the Caki‐1 cell line in order to determine their IC50 values. (Bromo)(NHC)silver(I) complexes 2a – 2c and (chloro)(NHC)gold(I) complexes 3a – 3c were found to have IC50 values of 27±2, 28±2, 34±6, 10±1, 12±5, and 12±3 μM , respectively, on the Caki‐1 cell line.  相似文献   

6.
Organotin(IV) complexes are known for their outstanding structural diversity and applications. Organotin(IV) carboxylates and dithiocarboxylates form an important class of organotin(IV) compounds. The structural diversity of these compounds emanates from several features including flexibility in coordination geometries, coordination numbers, and versatility of the ligands to engage in different modes of chelation from monodentate to bidentate. Triorganotin(IV) complexes with various ligands mostly demonstrate tetrahedral or trigonal bipyramidal symmetry with some distortions, while diorganotin(IV) and chlorodiorganotin(IV) complexes have variation of geometries and coordination numbers. Some monoorganotin(IV) complexes have also been reported with pentagonal bipyramidal geometries.  相似文献   

7.
The efficient one‐pot syntheses of Betti bases by the three‐component reaction of aromatic aldehyde, 2‐naphthalen, and acetonitrile (or benzamide) catalyzed by 1‐methyl‐3‐(2‐(sulfooxy)ethyl)‐1H‐imidazol‐3‐ium chloride is reported. The solvent can be recycled easily.  相似文献   

8.
The product from the reduction of ethyl spiro-3,3-(ethylenedioxy)-2-oxindole carboxylates (1) using borohydride salts has been found to be dependant upon both solvent and metal ion. With polar solvents and lithium bromide/sodium borohydride, spiro-3,3-(ethylenedioxy)-2-hydroxyindole carboxylates (2) are obtained in high yields whilst [2-(2-hydroxymethyl-[1,3]dioxolan-2-yl)-phenyl]-carbamic acid ethyl esters (3) are obtained using sodium borohydride in less polar solvents.  相似文献   

9.
Four zwitterions were prepared by treating 1,3‐dimesitylimidazolin‐2‐ylidene (SIMes) or 1,3‐dimesitylimidazol‐2‐ylidene (IMes) with either N‐tosyl benzaldimine or diphenylketene. They were isolated in high yields and characterized by IR and NMR spectroscopy. The molecular structures of three of them were determined by using X‐ray crystallography and their thermal stability was monitored by using thermogravimetric analysis. The imidazol(in)ium‐2‐amides were rather labile white solids that did not show any tendency to tautomerize into the corresponding 1,2,2‐triaminoethene derivatives. They displayed a mediocre catalytic activity in the Staudinger reaction of N‐tosyl benzaldimine with diphenylketene. In contrast, the imidazol(in)ium‐2‐enolates were orange‐red crystalline materials that remained stable over extended periods of time. Despite their greater stability, these zwitterions turned out to be efficient promoters for the model cycloaddition under scrutiny. As a matter of fact, their catalytic activity matched those recorded with the free carbenes. Altogether, these results provide strong experimental insight into the mechanism of the Staudinger reaction catalyzed by N‐heterocyclic carbenes. They also highlight the superior catalytic activity of the imidazole‐based carbene IMes compared with its saturated analogue SIMes in the reaction under consideration.  相似文献   

10.
The water exchange reaction of [Be(H2O)2(1H‐imidazole‐4,5‐dicarboxylate)] and [Be(H2O)2(1H‐imidazol‐3‐ium‐4,5‐dicarboxylate)]+ in water was studied by DFT calculations (RB3LYP/6‐311+G**) and identified as an associative interchange mechanism. The activation barriers for [Be(H2O)2(1H‐imidazole‐4,5‐dicarboxylate)] (16.6 kcal/mol) and [Be(H2O)2(1H‐imidazol‐3‐ium‐4,5‐dicarboxylate)]+ (13.8 kcal/mol) are similar to the barrier for [Be(H2O)4)]2+ and independent of the overall charge. NICS calculations show no indication that the aromaticity of the imidazole ring is affected during the water exchange process.  相似文献   

11.
Two new 1‐D silver( I ) complexes, [Ag( I )1.5(IM‐oBA)(NO3)0.5] ( 1 ) and [Ag( I )(NIT‐pBA)] ( 2 ), (IM‐oBAH = 2‐(2‐carboxyphenyl)‐4,4,5,5‐tetramethyl‐4,5‐dihydro‐1H‐imidazol‐1‐yloxyl, NIT‐pBAH = 2‐(4‐carboxyphenyl)‐4,4,5,5‐tetramethyl‐4,5‐dihydro‐3‐oxido‐H‐imidazol‐3‐ium‐1‐yloxyl) have been prepared and structurally characterized. Complexes 1 and 2 crystallized in the monoclinic space groups of C2/c and P21/c, respectively. In complex 1 , the structure consists of trinuclear Ag(I) atoms with different linkages of IM‐oBA and nitrates. The trinuclear Ag(I) atoms are further coordinated to the neighbor IM‐oBA radicals via self‐assembly of the nitrogen atom of imine moiety and extended into formation of a polymeric chain. Complex 2 is constructed from a bis(carboxylato‐O,O')‐bridged centrosymmetric dimeric subunit and extended into a polymeric chain through self‐assembly coordination between metal ions and nitroxide groups of NIT‐pBA radicals. Temperature dependence of magnetic susceptibility measurements showed a weak ferromagnetic coupling between nitroxide radicals in 1 and 2 with J = 5.62 for 1 and 6.62 cm?1 for 2 , respectively.  相似文献   

12.
The synthesis and characterization of two new complexes (IPr)Pd(acac)2 (1) and (IPr)Pd(acac)Cl (2) (IPr=(N,N'-bis(2,6-diisopropylphenyl)imidazol)-2-ylidene, acac=acetylacetonate) are described. Complex 2 can be prepared in a one-pot protocol in high yield. A study detailing the versatility of 2 to effectively catalyze a series of cross-coupling reactions is discussed.  相似文献   

13.
Two polyoxometallate‐based compounds, tris[1,1′‐(butane‐1,4‐diyl)bis(1H‐imidazol‐3‐ium)] bis[tetracosa‐μ2‐oxido‐dodecaoxido‐μ12‐phosphato‐dodecamolybdenum(VI)], (C10H16N4)3[PMo12O40]2, (I), and 1,1′‐(butane‐1,4‐diyl)bis(1H‐imidazol‐3‐ium) 1‐[4‐(1H‐imidazol‐1‐yl)butyl]‐1H‐imidazol‐3‐ium tetracosa‐μ2‐oxido‐dodecaoxido‐μ12‐phosphato‐dodecamolybdenum(VI) dihydrate, (C10H16N4)(C10H15N4)[PMo12O40]·2H2O, (II), were synthesized by hydrothermal techniques at different pH values. The stoichiometric ratio between the polyoxometallate (POM) anions and organic cations is 2:3 in (I), with one of the cations lying on an inversion centre. The doubly protonated 1,1′‐(butane‐1,4‐diyl)diimidazole (BIM) cations are linked to the [PMo12O40]3− anions by hydrogen bonds to form a three‐dimensional supramolecular network. The stoichiometric ratio of POM anions and organic cations is 1:2 in (II), and the anion is located about a centre of inversion. The partly protonated BIM cations and solvent water molecules form hydrogen bonds with the [PMo12O40]3− anions, yielding a two‐dimensional supramolecular layer. The different lattice architectures of (I) and (II) may be governed by the ratio between the POM anions and organic cations, which, in turn, is determined by the pH value.  相似文献   

14.
The present paper reports the structures of bis(adeninium) zoledronate tetrahydrate {systematic name: bis(6‐amino‐7H‐purin‐1‐ium) hydrogen [1‐hydroxy‐2‐(1H‐imidazol‐3‐ium‐1‐yl)‐1‐phosphonatoethyl]phosphonate tetrahydrate}, 2C5H6N5+·C5H8N2O7P22−·4H2O, (I), and bis(adeninium) zoledronate hexahydrate {systematic name: a 1:1 cocrystal of bis(6‐amino‐7H‐purin‐1‐ium) hydrogen [1‐hydroxy‐2‐(1H‐imidazol‐3‐ium‐1‐yl)‐1‐phosphonatoethyl]phosphonate hexahydrate and 6‐amino‐7H‐purin‐1‐ium 6‐amino‐7H‐purine dihydrogen [1‐hydroxy‐2‐(1H‐imidazol‐3‐ium‐1‐yl)ethane‐1,1‐diyl]diphosphonate hexahydrate}, 2C5H6N5+·C5H8N2O7P22−·6H2O, (II). One of the adenine molecules and one of the phosphonate groups of the zoledronate anion of (II) are protonated on a 50% basis. The zoledronate group displays its usual zwitterionic character, with a protonated imidazole ring; however, the ionization state of the phosphonate groups of the anion for (I) and (II) are different. In (I), the anion has both singly and doubly deprotonated phosphonate groups, while in (II), it has one singly deprotonated phosphonate group and a partially deprotonated phosphonate group. In (I), the cations form an R22(10) base pair, while in (II), they form R22(8) and R22(10) base pairs. Two water molecules in (I) and five water molecules in (II) are involved in water–water interactions. The presence of an additional two water molecules in the structure of (II) might influence the different ionization state of the anion as well as the different packing mode compared to (I).  相似文献   

15.
Carbon nucleophiles generated by a non-nucleophilic base (LDA) were effectively trapped with di-tert-butyl dicarbonate (Boc-anhydride) to provide the corresponding tert-butyl aryl acetates, di-tert-butyl aryl malonates, unsymmetrical aryl malonates and tert-butyl benzoates in high yields. This reaction represents another useful way to prepare a variety of tert-butyl carboxylates and highlights the synthetic utility of di-tert-butyl dicarbonate as a versatile carboxylating reagent.  相似文献   

16.
An efficient synthesis of 7‐alkyl‐6H,7H‐naphtho‐[10,20:5,6]pyrano[3,2‐c]chromen‐6‐ones by three‐component condensation reaction of β‐naphthol, aromatic aldehydes, and 4‐hydroxycoumarin catalyzed by 1‐methyl‐3‐(2‐(sulfooxy)ethyl)‐1H‐imidazol‐3‐ium chloride is reported in good to excellent yields and short reaction times.  相似文献   

17.
An amino group has been introduced into one ligand of DOTA that can couple to peptidyl carboxylates by coupling α-brominated glycine to DO3A-tBu (1,4,7,10-tetraazacyclododecane-1,4,7-tris(acetic acid, tert-butylester)). α-Amino-DOTA was coupled to the carboxylate backbone terminus of a peptide to demonstrate the utility for derivatization.  相似文献   

18.
The title compound, C6H9N2O2+·Cl·C6H8N2O2·H2O, contains one 2‐(3‐methyl‐1H‐imidazol‐3‐ium‐1‐yl)acetate inner salt molecule, one 1‐carboxymethyl‐3‐methyl‐1H‐imidazol‐3‐ium cation, one chloride ion and one water molecule. In the extended structure, chloride anions and water molecules are linked via O—H...Cl hydrogen bonds, forming an infinite one‐dimensional chain. The chloride anions are also linked by two weak C—H...Cl interactions to neighbouring methylene groups and imidazole rings. Two imidazolium moieties form a homoconjugated cation through a strong and asymmetric O—H...O hydrogen bond of 2.472 (2) Å. The IR spectrum shows a continuous D‐type absorption in the region below 1300 cm−1 and is different to that of 1‐carboxymethyl‐3‐methylimidazolium chloride [Xuan, Wang & Xue (2012). Spectrochim. Acta Part A, 96 , 436–443].  相似文献   

19.
By the adoption of annulated ring systems for their steric bulkiness, a new series of symmetric N ‐heterocyclic imidazolium and perimidium salts were synthesized. Also, corresponding asymmetric ferrocenyl N ‐functionalized series were prepared as salts. All the reported salts were fully characterized. The reported X‐ray structure of 4,5‐diphenyl‐1,3‐dimethyl‐1H ‐imidazol‐3‐ium iodide ( 2a ) shows that it crystallized in the orthorhombic space group P 212121. Low to moderate antimicrobial activities were observed with both sets of salts against important clinical isolates of Staphylococcus aureus , Bacillus subtilis , and Enterococcus faecalis and Escherichia coli , Pseudomonas aeruginosa , and Salmonella enterica . The results were benchmarked against meropenem. Both 1,3‐propyl‐1H ‐phenanthro[9,10‐d ]imidazol‐3‐ium iodide ( 3b ) and 1‐ferrocenyl‐3‐propyl‐1H ‐perimidin‐3‐ium iodide ( 9b ) showed high antimicrobial activities against all tested Gram‐positive bacterial strains, with minimum inhibitory concentration values ranging from 8 to 4 μg/mL (meropenem = 0.5–0.125 μg/mL), while all the salts showed little or no activity against Gram‐negative bacterial strains. In general, the asymmetric ferrocenyl‐containing salts exhibited higher activities than the symmetric ones.  相似文献   

20.
Treatment of a mixture of alkyl aldehydes (1) with acid chlorides (2) in the presence of zinc metal powder and a catalytic amount of chlorotrimethylsilane (TMSCl) in dichloromethane brought about highly facile and effective coupling to give selectively the corresponding α-haloacylation and gem-bisacylation products, α-haloalkyl carboxylates and 1,1-dicarboxylates (acylals), in good to excellent yields.  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