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.
何永炳  LEHN  Jean-Marie 《中国化学》2000,18(3):384-387
The 6-hydroxymethyl-6'-tetrahydropyranyloxymethyl-2, 2'-bipyridine (2) was synthesized by the reaction of 6,6'-dihy-droxymethyl-2, 2'-bipyridine (1) with 3, 4-dihydropyran (DHP). 6-Tetrahydropyranyloxymethyl-6'-iodomethyl-2, 2'-bipyridine (5) was obtained from mesylate and iodizating reaction of compound 2. The coupling of 2 and 5 followed by hydrolysis gave bis(6'-hydroxymethyl-2,2'-bipyridme-6-methyl)ether (7) .The macrocydic ligand 8 was obtained by treating 7 and 6,6'-dibromomethyl-2,2'-bipyridine. The synthetic conditions of the intermediate 2 and macocyclic ligand 8 were discussed.  相似文献   

2.
Yoon I  Seo J  Lee JE  Park KM  Kim JS  Lah MS  Lee SS 《Inorganic chemistry》2006,45(9):3487-3489
The S3O2 macrocycle L1 was synthesized by a dithiol-dihalide coupling reaction under high-dilution conditions. The reaction of L1 with K2PdCl4 afforded an exocoordinated complex 1, [cis-Cl2Pd(L1)], which can then be manipulated to provide a heterobinuclear complex 3, {[Pd(L1)Ag(NO3)(2.5)](NO3)(0.5)}n, utilizing endocyclic Pd(II) and exocyclic Ag(I) in a single macrocycle through a successive reaction with AgNO3. The network of 3 contains a unique honeycomb-like 2-D sheet made up of the repeating unit [Ag6(NO3)6].  相似文献   

3.
Density functional calculations are employed to theoretically explore the mechanism of all elementary reaction steps involved in the catalytic reaction of 6-phosphogluconate dehydrogenase (6PGDH). The model systems we choose for the enzyme contain the essential parts of the cofactor (NADP+), the substrate 6-phosphogluconate (6PG), and some key residues (Lys183 and Glu190) in the active site of sheep liver 6PGDH. The effect of the apoenzyme electrostatic environment on the studied reaction is treated by the self-consistent reaction-field method. Our calculations demonstrate that the first step of the catalytic reaction is the formation of a 3-keto 6PG intermediate, which proceeds through a concerted transition state involving a hydride transfer from 6PG to NADP+, and a proton transfer from 6PG to Lys183. The second step is the elimination of a CO2 molecule from 6-PG, concomitant with a proton transfer from Lys183 to 6-PG. In the final step, a concerted double proton transfer (one from Glu190 to the substrate, another from the substrate to Lys183) results in the final product, the keto form of ribulose 5-phosphate (Ru5P). The rate-limiting step is the formation of a 3-keto 6PG intermediate, with a free energy barrier of 22.7 kcal/mol at room temperature in the protein environment, and all three steps are calculated to be thermodynamically favorable. These results are in good agreement with the general acid/general base mechanism suggested from previous experiments for the 6PGDH reaction.  相似文献   

4.
The reaction of the ketenyl radical (HCCO) with acetylene (C(2)H(2)) is very relevant to the oxygen-acetylene flames and fuel-rich combustion process for nitrogen-containing compounds. Unfortunately, except for several rate constant measurements, the mechanism is completely unknown for this reaction. In this paper, detailed theoretical investigations are performed for the HCCO + C(2)H(2) reaction at the G3B3 level using the B3LYP/6-31G(d), B3LYP/6-311++G(d,p), and QCISD/6-31G(d) geometries. The exclusive fragmentation channel is the formation of the cyclopropenyl radical (c-C(3)H(3)) and carbon monoxide (CO) via the chainlike OCCHCHCH and three-membered ring OC-cCHCHCH intermediates. Thus, the mass spectroscopic peak of C(3)H(3)(+) in a previous experiment can be explained. The calculated overall reaction barrier is 4.4, 4.4, and 5.3 kcal/mol at the G3B3//B3LYP/6-31G(d), G3B3//B3LYP/6-311++G(d,p), and G3B3//QCISD/6-31G(d) levels, respectively. The title reaction may provide an effective route for generating the long-sought cyclopropenyl radical in the laboratory, which has been the long-standing subject of numerous theoretical studies as the simplest cyclic conjugate radical, and its bulky derivatives were already known. Future experimental investigations for the HCCO + C(2)H(2) reaction are greatly desired to test the predicted fragmentation channel. The implication of the present study in combustion and interstellar processes is discussed.  相似文献   

5.
Zhou L  Wang J  Zhang Y  Yao Y  Shen Q 《Inorganic chemistry》2007,46(14):5763-5772
The synthesis and structures of a series of lanthanide(II) and lanthanide(III) complexes supported by the amido ligand N(SiMe3)Ar were described. Several lanthanide(III) amide chlorides were synthesized by a metathesis reaction of LnCl3 with lithium amide, including {[(C6H5)(Me3Si)N]2YbCl(THF)}2.PhCH3 (1), [(C6H3-iPr2-2,6)(SiMe3)N]2YbCl(mu-Cl)Li(THF)3.PhCH3 (4), [(C6H3-iPr2-2,6)(SiMe3)N]YbCl2(THF)3 (6), and [(C6H3-iPr2-2,6)(SiMe3)N]2SmCl3Li2(THF)4 (7). The reduction reaction of 1 with Na-K alloy afforded bisamide ytterbium(II) complex [(C6H5)(Me3Si)N]2Yb(DME)2 (2). The same reaction for Sm gave an insoluble black powder. An analogous samarium(II) complex [(C6H5)(Me3Si)N]2Sm(DME)2 (3) was prepared by the metathesis reaction of SmI2 with NaN(C6H5)(SiMe3). The reduction reaction of ytterbium chloride 4 with Na-K alloy afforded monoamide chloride {[(C6H3-iPr2-2,6)(SiMe3)N]Yb(mu-Cl)(THF)2}2 (5), which is the first example of ytterbium(II) amide chloride, formed via the cleavage of the Yb-N bond. The same reduction reaction of 7 gave a normal bisamide complex [(C6H3-iPr2-2,6)(SiMe3)N]2Sm(THF)2 (8) via Sm-Cl bond cleavage. This is the first example for the steric effect on the outcome of the reduction reaction in lanthanide(II) chemistry. 5 can also be synthesized by the Na/K alloy reduction reaction of 6. All of the complexes were fully characterized including X-ray diffraction for 1-7.  相似文献   

6.
利用原子转移自由基聚合(ATRP)合成了一种新型的含假芪型偶氮生色团的两亲性嵌段共聚物P(HEMA-b-6CNAzo)。首先,采用ATRP引发剂引发三甲基硅保护的羟乙基甲基丙烯酸酯(HEMA—TMS)聚合,得到大分子引发剂P(HEMA—TMS);接着进一步引发单体甲基丙烯酸6-(N_甲基苯胺基)己酯进行ATRP反应,得...  相似文献   

7.
硫代硫酸根插层水滑石的层间限域反应   总被引:1,自引:0,他引:1  
将无机阴离子硫代硫酸根(S2O23-)限域在锌铝水滑石(LDH)层间,并研究了其在水滑石层板限域空间内被铁氰根(Fe(CN)63-)氧化的反应过程.通过X射线衍射(XRD)和傅里叶变换红外(FTIR)光谱仪对反应的中间产物和最终产物进行的表征发现,氧化产物连四硫酸根(S4O62-)进入到溶液中,还原产物亚铁氰根(Fe(CN)64-)则保留在水滑石层间.进一步系统研究了该反应的动力学过程,考察了硫代硫酸根插层水滑石用量、铁氰化钾浓度和温度对反应的影响.结果表明该氧化还原反应符合球体内扩散模型.根据温度对反应速率影响,得出了该反应的表观活化能为24.6kJ.mol-1,比相同条件下溶液中反应活化能降低了约13.7kJ.mol-1.采用分子动力学(MD)模拟计算了水分子含量对硫代硫酸根插层水滑石层间距大小的影响.计算表明:在水溶液环境中,水滑石微反应器的尺寸在特定方向具有可调控性.根据实验表征和理论计算对该层间反应的机理进行了探讨.因此,该类层状材料可以作为一种新型纳米级微反应器应用于调控化学反应.  相似文献   

8.
Cysteine oxidation by HO(.) was studied at a high level of ab initio theory in both gas phase and aqueous solution. Potential energy surface scans in the gas phase performed for the model system methanethiol+HO(.) indicate that the reactants can form two intermediate states: a sulfur-oxygen adduct and a hydrogen bound reactant complex. However these states appear to play a minor role in the reaction mechanism as long as they are fast dissociating states. Thus the main reaction channel predicted at the QCISD(T)/6-311+G(2df,2pd) level of theory is the direct hydrogen atom abstraction. The reaction mechanism is not perturbed by solvation which was found to induce only small variations in the Gibbs free energy of different reactant configurations. The larger size reactant system cysteine+HO* was treated by the integrated molecular orbital+molecular orbital (IMOMO) hybrid method mixing the QCISD(T)/6-311+G(2df,2pd) and the UMP2/6-311+G(d,p) levels of theory. The calculated potential energy, enthalpy, and Gibbs free energy barriers are slightly different from those of methanethiol. The method gave a rate constant for cysteine oxidation in aqueous solution, k=2.4 x 10(9) mol(-1) dm(3) s(-1), which is in good agreement with the experimental rate constant. Further analysis showed that the reaction is not very sensitive to hydrogen bonding and electrical polarity of the molecular environment.  相似文献   

9.
研究了在阳离子表面活性剂存在下水/有机两相中水溶性铑配合物RhCI(CO)(TPPTS)2(TPPTS:P(m-C6H4SO3Na)3)催化双环戊二烯氢甲酰化反应,考察了反应温度、催化剂浓度、不同水溶性膦配体TPPTS和TPPDS(C5H5P(m-C6H4SO3Na)2),以及表面活性剂结构对催化反应的影响.结果表明,...  相似文献   

10.
This paper presents an ab initio (RHF/6-31G** and MP2(full)6-31G**) and density functional (DFT) study of the structure and energetics of formation of an intermolecular complex which is the simplest model of an active center lysozyme with a substrate. The calculated energy of complex formation is 41.4 (RHF), 53.4 (MP2), and 52.7 kcal/mole (DFT). The proton transfer reaction is a concerted reaction having an energy barrier of 41.1 (RHF), 31.6 (MP2), and 25.3 (DFT) kcal/mole.  相似文献   

11.
The reaction of [RuCp(IPri)(CH3CN)2]PF6 (IPri = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) with HCCR (R = COOMe, COOEt, COMe) yields the allyl carbene complexes [RuCp(=C(R)-eta3-CHC(R)CH-IPri)]PF6. This conversion involves selective head-to-tail coupling of two alkynes and an unusual migratory insertion of the N-heterocyclic carbene into the ruthenium-carbon double bond of a ruthenacyclopentatriene intermediate.  相似文献   

12.
Zhang B  Zhang L  Sun L  Cui Z 《Organic letters》2002,4(21):3615-3618
[reaction: see text] The trinucleotide cytidylyl(3'-->5'phosphoryl)cytidylyl(3'-->5'phosphoryl)-3'-deoxy-3'-(L-phenylalanyl) amido adenosine (CpCpA-NH-Phe) was synthesized by phosphoramidite chemistry from 3'-amino-3'-deoxyadenosine as the ribosomal substrate. The 3'-amino-3'-deoxyadenosine was first converted to 3'-(N-tert-butyloxycarbonyl-L-phenylalanine)amido-3'-deoxy-6-N,6-N,2'-O-tribenzoyl-adenosine and then coupled with cytidine phosphoramidite to produce the fully protected CpCpA-NH-Phe-Boc. The title product was obtained after removing all protection groups and then radiolabeled with (32)P to yield pCpCpA-NH-Phe, which demonstrated high activity for the peptidyl transferase reaction in the ribosome.  相似文献   

13.
The reactivity of the silylsilylene [{PhC(NtBu)(2)}SiSi(Cl){(NtBu)(2)C(H)Ph}] (2) towards diphenylacetylene, azobenzene, 2,6-diisopropylphenyl azide, sulfur, and selenium is described. The reaction of 2 with one equivalent of azobenzene in toluene afforded compound 3, which is the first example of a 1,2-diaza-3,4-disilacyclobutane containing a pentacoordinate silicon center. The formation of 3 can be explained by a [1+2] cycloaddition of the divalent Si center in 2 with PhN=NPh to form a diazasilacyclopropane intermediate, which then undergoes a 1,2-chlorine shift to release the ring strain to form 3. Similarly, the reaction of 2 with one equivalent of diphenylacetylene in toluene afforded the 1,2-disilacyclobutene 4, which contains a pentacoordinate silicon center. The reaction of 2 with 1.6 equivalents of 2,6-diisopropylphenylazide in toluene afforded the silaimine [LSi(=NAr)N(Ar)L'] (5, L=PhC(NtBu)(2) , L'=Si(Cl){(NtBu)(2)C(H)Ph}, Ar=2,6-iPr(2)C(6)H(3)). The formation of 5 can be explained by an oxidative addition of the divalent Si center in 2 with ArN(3) to afford a silaimine intermediate, which then reacts with another molecule of ArN(3) to give compound 5. The reaction of 2 with elemental sulfur in toluene afforded the chlorosilanethione [LSi(S)Cl] (6) and dithiodisiletane [{Ph(H)C(NtBu)(2) }Si(μ-S)](2) (7). Treatment of 2 with elemental selenium in THF afforded the di(silaneselone) [LSi(Se)Si(Se)L] (8). Evidently, the divalent Si center in 2 undergoes oxidative addition with chalcogens to afford a silylsilanechalcogenone intermediate, which then displaces ":Si{(NtBu)(2)C(H)Ph}" and "ClSi{(NtBu)(2) C(H)Ph}" to form 6 and 8, respectively. Moreover, compound 8 was synthesized by the reaction of [{PhC(NtBu)(2)}Si:](2) (10) with elemental selenium in THF. The results show that the reactions of 2 are initiated by oxidative addition of the divalent silicon center, and then the intermediate formed undergoes a rearrangement involving the diaminochlorosilyl substituent to form compounds 3-8. These products have been characterized by NMR spectroscopy and X-ray crystallography.  相似文献   

14.
A dramatic enhancement of the diastereo- and enantioselectivity in the nitro-Michael addition reaction organocatalysed by a commercially available α,α-L-diaryl prolinol was disclosed when performing the reaction in unconventional hexafluorobenzene as a medium. DFT calculations were performed to clarify the origin of stereoselectivity and the role of C(6)F(6).  相似文献   

15.
This paper reports a new method for the generation of chiral Lewis superacids by protonation of a non-Lewis acidic oxazaborolidine (1) with triflic acid. The resulting cationic species (3) are powerful and highly enantioselective catalysts for the Diels-Alder reaction of various 1,3-dienes with alpha,beta-enals. An optimization study (see Table 1) led to the selection of reaction conditions and catalysts (6A and 6B) which are very effective. The reactions are simple to conduct, reproducible, and economical, since only ca. 6 mol % of catalyst is required. In addition, the chiral catalyst precursor is readily recovered for reuse (>95% efficiency) and is commercially available. The broad scope of the process is documented by the 14 examples listed in Table 2. The absolute stereochemical course of the Diels-Alder reactions catalyzed by 6A and 6B was successfully predicted on the basis of the mechanistic principles which have recently been formulated for this type of catalytic enantioselective reaction involving re-face attack by the diene on complex 7. The mode of generation of Lewis superacids 6A and 6B allows an approximate comparison (or scale) connecting the catalytic power Lewis and protic acids.  相似文献   

16.
The reaction of 2,2'-bisdipyrrins 1, 2, and 3 with manganese(II) acetate tetrahydrate and molecular dioxygen yields the manganese(III)corroles 4, 5 and 6, which are readily demetalated to the respective free-base corroles.  相似文献   

17.
Ab initio calculations were carried out for the reaction of adamantylideneadamantane (Ad=Ad) with Br2 and 2Br2. Geometries of the reactants, transition states, intermediates, and products were optimized at HF and B3LYP levels of theory using the 6-31G(d) basis set. Energies were also obtained using single point calculations at the MP2/6-31G(d)//HF/6-31G(d), MP2/6-31G(d)//B3LYP/6-31G(d), and B3LYP/6-31+G(d)//B3LYP/6-31G(d) levels of theory. Intrinsic reaction coordinate (IRC) calculations were performed to characterize the transition states on the potential energy surface. Only one pathway was found for the reaction of Ad=Ad with one Br2 producing a bromonium/bromide ion pair. Three mechanisms for the reaction of Ad=Ad with 2Br2 were found, leading to three different structural forms of the bromonium/Br3- ion pair. Activation energies, free energies, and enthalpies of activation along with the relative stability of products for each reaction pathway were calculated. The reaction of Ad=Ad with 2Br2 was strongly favored over the reaction with only one Br2. According to B3LYP/6-31G(d) and single point calculations at MP2, the most stable bromonium/Br3- ion pair would form spontaneously. The most stable of the three bromonium/Br3- ion pairs has a structure very similar to the observed X-ray structure. Free energies of activation and relative stabilities of reactants and products in CCl4 and CH2ClCH2Cl were also calculated with PCM using the united atom (UA0) cavity model and, in general, results similar to the gas phase were obtained. An optimized structure for the trans-1,2-dibromo product was also found at all levels of theory both in gas phase and in solution, but no transition state leading to the trans-1,2-dibromo product was obtained.  相似文献   

18.
The detailed reaction profiles of the neutral-neutral as well as the cation-neutral direct hydroamination reactions between ethylene and ammonia are analyzed using MP2 (Full)/6-31++G(2df,2p) and B3LYP/6-31++G(2df,2p) methodologies. Analysis shows that both neutral-neutral, as well as the cation-neutral reactions are exothermic and the latter is >100 kJ/mol more exothermic than the former. Calculations show that a very large barrier height (>200 kJ/mol), and very large negative reaction entropy prevent the neutral-neutral reaction from proceeding in the forward direction. Analysis of the cation-neutral reaction, which is barrierless (the transition state is more stable than the reactants) and highly exothermic, indicates that the direct hydroamination reaction is thermodynamically attainable via a cation-neutral reaction pathway without a catalyst. Our calculations also suggest that although the cation-neutral direct hydroamination reaction is very fast, the cation of either ethylene or ammonia goes through a structural relaxation process before reacting with the other neutral reactant.  相似文献   

19.
Diphenylprolinol silyl ether was found to be an effective organocatalyst for promoting the asymmetric, catalytic, intramolecular [6 + 2] cycloaddition reactions of fulvenes substituted at the exocyclic 6-position with a δ-formylalkyl group to afford synthetically useful linear triquinane derivatives in good yields and excellent enantioselectivities. The cis-fused triquinane derivatives were obtained exclusively; the trans-fused isomers were not detected among the reaction products. The intramolecular [6 + 2] cycloaddition occurs between the fulvene functionality (6π) and the enamine double bond (2π) generated from the formyl group in the substrates and the diphenylprolinol silyl ether. The absolute configuration of the reaction products was determined by vibrational circular dichroism. The reaction mechanism was investigated using molecular orbital calculations, B3LYP and MP2 geometry optimizations, and subsequent single-point energy evaluations on model reaction sequences. These calculations revealed the following: (i) The intermolecular [6 + 2] cycloaddition of a fulvene and an enamine double bond proceeds in a stepwise mechanism via a zwitterionic intermediate. (ii) On the other hand, the intramolecular [6 + 2] cycloaddition leading to the cis-fused triquinane skeleton proceeds in a concerted mechanism via a highly asynchronous transition state. (iii) The fulvene functionality and the enamine double bond adopt the gauche-syn conformation during the C-C bond formation processes in the [6 + 2] cycloaddition. (iv) The energy profiles calculated for the intramolecular reaction explain the observed exclusive formation of the cis-fused triquinane derivatives in the [6 + 2] cycloaddition reactions. The reasons for the enantioselectivity seen in these [6 + 2] cycloaddition reactions are also discussed.  相似文献   

20.
The present paper contains the reaction mechanism of the rare earth complex with 2-(2-arsenophenylazo)-7-(2, 6-dichloro-4-fluoro-phenylazo)- 1, 8-dihydroxynaphthalene-3, 6-disulfonic acid (DCF-arsenazo) studied preliminarily with a fully automated stopped-flow analyzer. A method with a high sensitivity and selectivity is proposed for the simultaneous determination of samarium and gadolinium based on the differential kinetic reaction of the rare earth complex with DCF-arsenazo. Sm and Gd in their concentrated oxides were determined with the relative errors less than 6%.  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