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.
The first zwitterionic borata‐bis(NHC)‐stabilized phosphaketenyl germyliumylidene [(L2(O=C=P)Ge:] 2 (L2=(p ‐tolyl)2B[1‐(1‐adamantyl)‐3‐yl‐2‐ylidene]2) has been synthesized by salt‐metathesis reaction of [L2(Cl)Ge:] 1 with sodium phosphaethynolate [(dioxane)n NaOCP]. Unexpectedly, its exposure to UV light affords, after reductive elimination of the entire PCO group, the unprecedented [L2Ge‐GeL2] complex 3 in 54 % yields bearing the Ge22+ ion with Ge in the oxidation state +1. In addition, the 1,3‐digermylium‐2,4‐diphosphacyclobutadiene [L2Ge(μ‐P)2GeL2] 4 and bis(germyliumylidenyl)‐substituted diphosphene [(L2Ge‐P=P‐GeL2)] 5 could also be obtained in moderate yields. The formation of 3 – 5 and their electronic structures have been elucidated with DFT calculations.  相似文献   

2.
Reactivity studies of the thermally stable ruthenostannylene complex [Cp*(IXy)(H)2Ru? Sn? Trip] ( 1 ; IXy=1,3‐bis(2,6‐dimethylphenyl)imidazol‐2‐ylidene; Cp*=η5‐C5Me5; Trip=2,4,6‐iPr3C6H2) with a variety of organic substrates are described. Complex 1 reacts with benzoin and an α,β‐unsaturated ketone to undergo [1+4] cycloaddition reactions and afford [Cp*(IXy)(H)2RuSn(κ2‐O,O‐OCPhCPhO)Trip] ( 2 ) and [Cp*(IXy)(H)2RuSn(κ2‐O,C‐OCPhCHCHPh)Trip] ( 3 ), respectively. The reaction of 1 with ethyl diazoacetate resulted in a tin‐substituted ketene complex [Cp*(IXy)(H)2RuSn(OC2H5)(CHCO)Trip] ( 4 ), which is most likely a decomposition product from the putative ruthenium‐substituted stannene complex. The isolation of a ruthenium‐substituted stannene [Cp*(IXy)(H)2RuSn(?Flu)Trip] ( 5 ) and stanna‐imine [Cp*(IXy)(H)2RuSn(κ2‐N,O‐NSO2C6H4Me)Trip] ( 6 ) complexes was achieved by treatment of 1 with 9‐diazofluorene and tosyl azide, respectively.  相似文献   

3.
Reactivity studies of the thermally stable ruthenostannylene complex [Cp*(IXy)(H)2Ru Sn Trip] ( 1 ; IXy=1,3‐bis(2,6‐dimethylphenyl)imidazol‐2‐ylidene; Cp*=η5‐C5Me5; Trip=2,4,6‐iPr3C6H2) with a variety of organic substrates are described. Complex 1 reacts with benzoin and an α,β‐unsaturated ketone to undergo [1+4] cycloaddition reactions and afford [Cp*(IXy)(H)2RuSn(κ2‐O,O‐OCPhCPhO)Trip] ( 2 ) and [Cp*(IXy)(H)2RuSn(κ2‐O,C‐OCPhCHCHPh)Trip] ( 3 ), respectively. The reaction of 1 with ethyl diazoacetate resulted in a tin‐substituted ketene complex [Cp*(IXy)(H)2RuSn(OC2H5)(CHCO)Trip] ( 4 ), which is most likely a decomposition product from the putative ruthenium‐substituted stannene complex. The isolation of a ruthenium‐substituted stannene [Cp*(IXy)(H)2RuSn(Flu)Trip] ( 5 ) and stanna‐imine [Cp*(IXy)(H)2RuSn(κ2‐N,O‐NSO2C6H4Me)Trip] ( 6 ) complexes was achieved by treatment of 1 with 9‐diazofluorene and tosyl azide, respectively.  相似文献   

4.
The sterically demanding β‐diketiminate ligand Ldmp [Ldmp = HC{(CMe)N(dmp)}2, dmp = C6H3‐2,6‐Me2] was used to stabilize various gallium complexes in the formal oxidation states +II and +III. The reaction of in situ generated [LdmpLi] with gallium chloride affords [LdmpGaCl2] ( 1 ), which was used as starting complex to synthesize a variety of gallium(III) compounds [LdmpGaX2] [X = F ( 2 ), I ( 3 ), H ( 4 ), and Me ( 5 )]. Synthesis of the dinuclear complex [LdmpGaI]2 ( 6 ), with gallium in the formal oxidation state +II was accomplished by converting “GaI” with in situ generated [LdmpLi] in toluene. All compounds were characterized by elemental analyses, NMR spectroscopy, LIFDI‐TOF‐MS, and single‐crystal X‐ray diffraction. Additionally DFT calculations were performed for analysis of the bonding in 6 .  相似文献   

5.
郭倩玲  马淑兰  朱文祥  刘迎春  张静 《中国化学》2005,23(10):1387-1390
The X-ray crystallographic structure was reported for a dinuclear copper(Ⅱ) complex with a tetraanionic ligand of p-tert-butylsulfonylcalix[4]arene [Cu2L(CH3OH)6]·4CH3OH (H4L=p-tert-butylsuffonylcalix[4]arene). The complex belongs to triclinic system, P1^-- space group, with a = 1.2303(3) nm, b = 1.2377(3) nm, c = 1.3110(3) nm, a =66.862(4)°, β= 67.206(4)°, γ=61.711(3)°, Z= 1, V= 1.5659(7) nm^3, Dc= 1.371 g/cm^3, F(000) = 682,μ(Mo Kα) = 0.883 mm^-1, R1 =0.0325, wR2=0.0870. In this complex, the calix[4]arene acts as a bis-tridentate chelating ligand with the 1,2-alternate conformation.  相似文献   

6.
The new ruthenium complex [Ru(N3P)(OAc)][BPh4] ( 4 ), in which N3P is the N,P mixed tetradentate ligand N,N‐bis[(pyridin‐2‐yl)methyl]‐[2‐(diphenylphosphino)phenyl]methanamine was synthesized. The complex was found to be catalytically active for the endo cycloisomerization of alkynols. The catalytic reactions can be used to synthesize five‐, six‐, and seven‐membered endo‐cyclic enol ethers in good to excellent yields. A catalytic cycle involving a vinylidene intermediate was proposed for the catalytic reactions. Treatment of complex 4 with PhC?CH and H2O gave the alkyl complex [Ru(CH2Ph)(CO)(N3P)][BPh4] ( 30 ), which supports the assumption that the catalytic reactions involve addition of a hydroxyl group to the C?C bond of vinylidene ligands.  相似文献   

7.
Reactions of [Ru]Cl ([Ru]={Cp(PPh3)2Ru}; Cp=cyclopentadienyl) with three alkynyl compounds, 1 , 5 , and 8 , each containing a cyclobutyl group, are explored. For 1 , the reaction gives the vinylidene complex 2 , with a cyclobutylidene group, through dehydration at CδH and CγOH. With an additional methylene group, compound 5 reacts with [Ru]Cl to afford the cyclic oxacarbene complex 6 . The reaction proceeds via a vinylidene intermediate followed by an intramolecular cyclization reaction through nucleophilic addition of the hydroxy group onto Cα of the vinylidene ligand. Deprotonation of 2 with NaOMe produces the acetylide complex 3 and alkylations of 3 by allyl iodide, methyl iodide, and ethyl iodoacetate generate 4 a – c , respectively, each with a stable cyclobutyl group. Dehydration of 1 is catalyzed by the cationic ruthenium acetonitrile complex at 70 °C to form the 1,3‐enyne 7 . The epoxidation reaction of the double bond of 7 yields oxirane 8 . Ring expansion of the cyclobutyl group of 8 is readily induced by the acidic salt NH4PF6 to afford the 2‐ethynyl‐substituted cyclopentanone 9 . The same ring expansion is also seen in the reaction of [Ru]Cl with 8 in CH2Cl2, affording the vinylidene complex 10 , which can also be obtained from 9 and [Ru]Cl. However, in MeOH, the same reaction of [Ru]Cl with 8 affords the bicyclic oxacarbene complex 12 a through an additional cyclization reaction. Transformation of 10 into 12 a is readily achieved in MeOH/HBF4, but, in MeOH alone, acetylide complex 11 is produced from 10 . In the absence of MeOH, cyclization of 10 , induced by HBF4, is followed by fluorination to afford complex 13 . Crystal structures of 6 and 12 a ′ were determined by single‐crystal diffraction analysis.  相似文献   

8.
A viologen derivative carrying a benzimidazole group ( V-P-I 2+; viologen–phenylene–imidazole V-P-I ) can be dimerized in water using cucurbit[8]uril (CB[8]) in the form of a 2:2 complex resulting in a negative shift of the guest pKa, by more than 1 pH unit, contrasting with the positive pKa shift usually observed for CB-based complexes. Whereas 2:2 complex protonation is unclear by NMR, silver cations have been used for probing the accessibility of the imidazole groups of the 2:2 complexes. The protonation capacity of the buried imidazole groups is reduced, suggesting that CB[8] could trigger proton release upon 2:2 complex formation. The addition of CB[8] to a solution containing V-P- I 3+ indeed released protons as monitored by pH-metry and visualized by a coloured indicator. This property was used to induce a host/guest swapping, accompanied by a proton transfer, between V-P-I 3+ ⋅ CB[7] and a CB[8] complex of 1-methyl-4-(4-pyridyl)pyridinium. The origin of this negative pKa shift is proposed to stand in an ideal charge state, and in the position of the two pH-responsive fragments inside the two CB[8] which, alike residues engulfed in proteins, favour the deprotonated form of the guest molecules. Such proton release triggered by a recognition event is reminiscent of several biological processes and may open new avenues toward bioinspired enzyme mimics catalyzing proton transfer or chemical reactions.  相似文献   

9.
A newly synthesized one‐dimensional (1D) hydrogen‐bonded (H‐bonded) rhodium(II)–η5‐semiquinone complex, [Cp*Rh(η5p‐HSQ‐Me4)]PF6 ([ 1 ]PF6; Cp*=1,2,3,4,5‐pentamethylcyclopentadienyl; HSQ=semiquinone) exhibits a paraelectric–antiferroelectric second‐order phase transition at 237.1 K. Neutron and X‐ray crystal structure analyses reveal that the H‐bonded proton is disordered over two sites in the room‐temperature (RT) phase. The phase transition would arise from this proton disorder together with rotation or libration of the Cp* ring and PF6? ion. The relative permittivity εb′ along the H‐bonded chains reaches relatively high values (ca., 130) in the RT phase. The temperature dependence of 13C CP/MAS NMR spectra demonstrates that the proton is dynamically disordered in the RT phase and that the proton exchange has already occurred in the low‐temperature (LT) phase. Rate constants for the proton exchange are estimated to be 10?4–10?6 s in the temperature range of 240–270 K. DFT calculations predict that the protonation/deprotonation of [ 1 ]+ leads to interesting hapticity changes of the semiquinone ligand accompanied by reduction/oxidation by the π‐bonded rhodium fragment, producing the stable η6‐hydroquinone complex, [Cp*Rh3+6p‐H2Q‐Me4)]2+ ([ 2 ]2+), and η4‐benzoquinone complex, [Cp*Rh+4p‐BQ‐Me4)] ([ 3 ]), respectively. Possible mechanisms leading to the dielectric response are discussed on the basis of the migration of the protonic solitons comprising of [ 2 ]2+ and [ 3 ], which would be generated in the H‐bonded chain.  相似文献   

10.
Reaction of p‐phenylenediacetonitrile (p‐phda) with AgCF3COO afforded the coordination polymer, {[Ag2(p‐phda)2] [Ag4(CF3COO)6]}n ( 1 ), where the 1D cationic [Ag2(p‐phda)2]2+ chain acts as host and the anionic [Ag4(CF3COO)6]2– as guest molecules occupy the channel between neighboring host chains. This is a rare crystal example of AgCF3COO complex consisting of cationic complex chains and anionic guests. In addition, complex 1 exhibits luminescence at room temperature in solid state.  相似文献   

11.
A mixed‐valence Mn complex {[MnIIMnIII(HL)2(4,4′‐bpy)(H2O)2] · (ClO4)(DMF)3(4,4′‐bpy)0.5}n ( 1 ) [H2L = 3‐(2‐phenol)‐5‐(pyridin‐2‐yl)‐1,2,4‐triazole] was synthesized and characterized by X‐ray single‐crystal structure analysis and magnetic susceptibility. Single‐crystal X‐ray analysis revealed that complex 1 has a dinuclear core, in which adjacent central MnIII atoms are linked by 4,4′‐bipyridine to form an infinite one‐dimensional (1D) molecular configuration. According to the Mn surrounding bond lengths and bond valence sum (BVS) calculations, we demonstrated that the Mn atom coordinated to the pyridine N atoms is in the +2 oxidation state, while another Mn atom coordinated to the phenolic oxygen atoms is in the +3 oxidation state. Magnetic susceptibility data of the complex 1 indicate that the ferromagnetic interaction dominates in this complex.  相似文献   

12.
Treatment of N,N‐chelated germylene [(iPr)2NB(N‐2,6‐Me2C6H3)2]Ge ( 1 ) with ferrocenyl alkynes containing carbonyl functionalities, FcC≡CC(O)R, resulted in [2+2+2] cyclization and formation of the respective ferrocenylated 3‐Fc‐4‐C(O)R‐1,2‐digermacyclobut‐3‐enes 2 – 4 [R = Me ( 2 ), OEt ( 3 ) and NMe2 ( 4 )] bearing intact carbonyl substituents. In contrast, the reaction between 1 and PhC(O)C≡CC(O)Ph led to activation of both C≡C and C=O bonds producing bicyclic compound containing two five‐membered 1‐germa‐2‐oxacyclopent‐3‐ene rings sharing one C–C bond, 4,8‐diphenyl‐3,7‐dioxa‐2,6‐digermabicyclo[3.3.0]octa‐4,8‐diene ( 5 ). With N‐methylmaleimide containing an analogous C(O)CH=CHC(O) fragment, germylene 1 reacted under [2+2+2] cyclization involving the C=C double bond, producing 1,2‐digermacyclobutane 6 with unchanged carbonyl moieties. Finally, 1 selectively added to the terminal double bond in allenes CH2=C=CRR′ giving rise to 3‐(=CRR′)‐1,2‐digermacyclobutanes [R/R′ = Me/Me ( 7 ), H/OMe ( 8 )] bearing an exo‐C=C double bond. All compounds were characterized by 1H, 13C{1H} NMR, IR and Raman spectroscopy and the molecular structures of 3 , 4 , 5 , and 8 were established by single‐crystal X‐ray diffraction analysis. The redox behavior of ferrocenylated derivatives 2 – 4 was studied by cyclic voltammetry.  相似文献   

13.
The regio‐ and absolute stereochemistry of (7S)‐N‐[4‐(3‐thienyl)tricyclo[4.2.1.02,5]non‐3‐en‐3‐ylcarbonyl]‐2,10‐camphorsultam tetrahydrofuran hemisolvate, C24H29NO3S2·0.5C4H8O, and (7S)‐N‐[4‐(4‐tolyl)tricyclo[4.2.1.02,5]non‐3‐en‐3‐ylcarbonyl]‐2,10‐camphorsultam, C27H33NO3S, have been established. One contains a half‐occupancy tetrahydrofuran solvent molecule located on a twofold axis and the other contains two crystallographically unique molecules which are nearly identical. The extended structures of both complexes can be explained via weak C—H...O interactions, which link the molecules together into two‐dimensional sheets in the ab plane for the thienyl complex and ultimately into a three‐dimensional structure for the tolyl derivative. The stereochemistry of both structures confirms that [2+2] cycloadditions of bicyclic alkenes and alkynes catalysed by ruthenium are exclusively exo.  相似文献   

14.
Two new metal‐organic coordination polymers[Eu(m‐BDC)1.5(MOPIP) · 1/2H2O]n ( 1 ) and [Co(m‐BDC)(MOPIP)2 · 2H2O]n ( 2 ) [m‐H2BDC = benzene‐1, 3‐dicarboxylic acid, MOPIP = 2‐(4‐methoxyphenyl)‐1H‐imidazo[4, 5‐f] 1 , 10 phenanthroline] were hydrothermally synthesized and structurally characterized by elemental analysis, IR spectroscopy, and single‐crystal X‐ray diffraction. The coordination polymers crystallize in monoclinic space group P21/m for 1 ( 2 : P21/n), with a = 9.779(2), b = 18.242(4), c = 17.146(3) Å, β = 106.41(3)° for 1 , and with a = 8.2153(16), b = 27.974(6), c = 17.974(4) Å, β = 100.40(3)° for 2 . The crystal structure of complex 1 is a zipper‐like chain of octacoordinate Eu3+ ions, in which Eu3+ ions are bridged in two coordination modes by m‐BDC2+ ligands and decorated by MOPIP ligands. The molecular structure of complex 2 consists of a hexacoordinte Co2+ atom, which generates a slightly distorted octahedral arrangement, and assembles into three‐dimensional supramolecular nets by π ··· π stacking interactions. Additionally, these two compounds show strong fluorescence in the solid state at room temperature. Natural bond orbital (NBO) analysis is performed by using the NBO method built in Gaussian 03 Program. The calculation results show a weak covalent interaction between the coordinated atoms and metal ions.  相似文献   

15.
合成了一种新型不对称Schiff碱铜前体配合物KCuL和一种化学组成为[(CuL)2Mn (H2O)2]·0.5CH3OH·0.5CH3OH的顺式异三核配合物,并通过元素分析、IR谱的方法对其进行了表征(其中H3L = N-(2-{[(1E)-(5-氯-2-羟基苯基)亚甲基]胺基}乙基)-2-羟基苯甲酰胺)。利用X-射线单晶衍射方法对三核配合物的晶体结构进行了测定。该三核配合物的每一晶胞单元含有一个顺式中性异三核分子和两个无序的甲醇分子。中心锰离子Mn2+处于O6形成的变形八面体几何构型,而两个配阴离子[CuL]-在Mn2+周围呈顺式排布。磁性表明该三核配合物不仅具有分子内反铁磁作用,而且三核单元之间具有弱的铁磁交换作用,磁参数分别为J = -12.1 cm-1, g = 2.20 and zj¢ = 1.37 cm-1.  相似文献   

16.
A neutral dinuclear vanadium complex containing both oxido and dioxidovanadium cores with hydrazone based ligand, [VO(OCH3)(CH3OH)(HL)VO2] ( 1 ) {H4L = bis[(E)‐N′‐(5‐bromo‐2‐hydroxybenzylidene)]‐carbohydrazide}, was synthesized and fully characterized by X‐ray crystallography and spectroscopic methods (IR, UV/Vis, NMR). The ligand acts as a trinegative hexadentate N3O3 donor ligand to form a dinuclear complex and during the reaction V4+ is oxidized to V5+. The coordination polyhedra are a VO5N distorted octahedron for the mono‐oxidovanadium core and a VO3N2 trigonal bipyramid for the dioxidovanadium core. The results of catalytic reactions indicate that 1 is a highly active catalyst in the clean epoxidation reaction of cis‐cyclooctene using aqueous hydrogen peroxide in acetonitrile. Cyclic voltammetric experiments of 1 in DMSO reveal two quasi‐reversible peaks due to the VO3+–VO2+ and VO2+–VO2 couples.  相似文献   

17.
The crystal structure and in vitro cytotoxicity of the amphiphilic ruthenium complex [ 3 ](PF6)2 are reported. Complex [ 3 ](PF6)2 contains a Ru?S bond that is stable in the dark in cell‐growing medium, but is photosensitive. Upon blue‐light irradiation, complex [ 3 ](PF6)2 releases the cholesterol–thioether ligand 2 and an aqua ruthenium complex [ 1 ](PF6)2. Although ligand 2 and complex [ 1 ](PF6)2 are by themselves not cytotoxic, complex [ 3 ](PF6)2 was unexpectedly found to be as cytotoxic as cisplatin in the dark, that is, with micromolar effective concentrations (EC50), against six human cancer cell lines (A375, A431, A549, MCF‐7, MDA‐MB‐231, and U87MG). Blue‐light irradiation (λ=450 nm, 6.3 J cm?2) had little influence on the cytotoxicity of [ 3 ](PF6)2 after 6 h of incubation time, but it increased the cytotoxicity of the complex by a factor 2 after longer (24 h) incubation. Exploring the unexpected biological activity of [ 3 ](PF6)2 in the dark elucidated an as‐yet unknown bifaceted mode of action that depended on concentration, and thus, on the aggregation state of the compound. At low concentration, it acts as a monomer, inserts into the membrane, and can deliver [ 1 ]2+ inside the cell upon blue‐light activation. At higher concentrations (>3–5 μm ), complex [ 3 ](PF6)2 forms supramolecular aggregates that induce non‐apoptotic cell death by permeabilizing cell membranes and extracting lipids and membrane proteins.  相似文献   

18.
Two chelate ligands for europium(III) having minocycline (=(4S,4aS,5aR,12aS)‐4,7‐bis(dimethylamino)‐1,4,4a,5,5a,6,11,12a‐octahydro‐3,10,12,12a‐tetrahydroxy‐1,11‐dioxonaphthacene‐2‐carboxamide; 5 ) as a VIS‐light‐absorbing group were synthesized as possible VIS‐light‐excitable stable Eu3+ complexes for protein labeling. The 9‐amino derivative 7 of minocycline was treated with H6TTHA (=triethylenetetraminehexaacetic acid=3,6,9,12‐tetrakis(carboxymethyl)‐3,6,9,12‐tetraazatetradecanedioic acid) or H5DTPA (=diethylenetriaminepentaacetic acid=N,N‐bis{2‐[bis(carboxymethyl)amino]ethyl}glycine) to link the polycarboxylic acids to minocycline. One of the Eu3+ chelates, [Eu3+(minocycline‐TTHA)] ( 13 ), is moderately luminescent in H2O by excitation at 395 nm, whereas [Eu3+(minocycline‐DTPA)] ( 9 ) was not luminescent by excitation at the same wavelength. The luminescence and the excitation spectra of [Eu3+(minocycline‐TTHA)] ( 13 ) showed that, different from other luminescent EuIII chelate complexes, the emission at 615 nm is caused via direct excitation of the Eu3+ ion, and the chelate ligand is not involved in the excitation of Eu3+. However, the ligand seems to act for the prevention of quenching of the Eu3+ emission by H2O. The fact that the excitation spectrum of [Eu3+(minocycline‐TTHA)] is almost identical with the absorption spectrum of Eu3+ aqua ion supports such an excitation mechanism. The high stability of the complexes of [Eu3+(minocycline‐DTPA)] ( 9 ) and [Eu3+(minocycline‐TTHA)] ( 13 ) was confirmed by UV‐absorption semi‐quantitative titrations of H4(minocycline‐DTPA) ( 8 ) and H5(minocycline‐TTHA) ( 12 ) with Eu3+. The titrations suggested also that an 1 : 1 ligand Eu3+ complex is formed from 12 , whereas an 1 : 2 complex was formed from 8 minocycline‐DTPA. The H5(minocycline‐TTHA) ( 12 ) was successfully conjugated to streptavidin (SA) (Scheme 5), and thus the applicability of the corresponding Eu3+ complex to label a protein was established.  相似文献   

19.
Two phosphine ligands of [Pd(PPh3)4] were substituted by π(C?S) coordination of 4‐bromodithiobenzoic acid methyl ester resulting in complex 1 . The same ester, after alkylation, afforded the dicationic complex bis(μ‐methanethiolato)tetrakis(triphenylphosphine)dipalladium(2+) bis(tetrafluoroborate) ( 2 ) from the same palladium source. A related thiolato‐bridged complex, bis(μ‐methanethiolato)bis(1‐methylpyridin‐2(1H)‐ylidene)bis(triphenylphosphine)dipalladium(2+) bis(tetrafluoroborate) ( 4 ) and the trinuclear cluster tris(μ‐methanethiolato)tris(triphenylphosphine)tripalladium(+)(3Pd? Pd) ( 5 ) resulted from treatment of a known cationic pyridinylidene complex with MeSLi. The double oxidative substitution reaction of [Pd(PPh3)4] with 1,5‐dichloro‐9,10‐anthraquinone afforded trans‐dichloro[μ‐(9,10‐dihydro‐9,10‐dioxoanthracene‐1,5‐diyl)]tetrakis(triphenylphosphine)dipalladium ( 6 ). Some of these complexes could be fully characterized by 1H‐, 13C‐, and 31P‐NMR spectroscopy, mass spectrometry, and elemental analysis. The crystal and molecular structures of all of them, and of trans‐bis(1,3‐dihydro‐1,3‐dimethyl‐2H‐imidazol‐2‐ylidene)diiodopalladium ( 3 ), were determined by single‐crystal X‐ray diffraction.  相似文献   

20.
Aldolases are enzymes that catalyze stereospecific aldol reactions in a reversible manner. Naturally occurring aldolases include class I aldolases, which catalyze aldol reactions via enamine intermediates, and class II aldolases, in which Zn2+ enolates of substrates react with acceptor aldehydes. In this work, Zn2+ complexes of L ‐prolyl‐pendant[15]aneN5 (ZnL3), L ‐prolyl‐pendant[12]aneN4 (ZnL4), and L ‐valyl‐pendant[12]aneN4 (ZnL5) were designed and synthesized for use as chiral catalysts for enantioselective aldol reactions. The complexation constants for L3 to L5 with Zn2+ [logKs(ZnL)] were determined to be 14.1 (for ZnL3), 7.6 (for ZnL4), and 9.6 (for ZnL5), indicating that ZnL3 is more stable than ZnL4 and ZnL5. The deprotonation constants of Zn2+‐bound water [pKa(ZnL) values] for ZnL3, ZnL4, and ZnL5 were calculated to be 9.2 (for ZnL3), 8.2 (for ZnL4), and 8.6 (for ZnL5), suggesting that the Zn2+ ions in ZnL3 is a less acidic Lewis acid than in ZnL4 and ZnL5. These values also indicated that the amino groups on the side chains weakly coordinate to Zn2+. We carried out aldol reactions between acetone and 2‐chlorobenzaldehyde and other aldehydes in the presence of catalytic amounts of the chiral Zn2+ complexes in acetone/H2O at 25 and 37 °C. Whereas ZnL3 yielded the aldol product in 43 % yield and 1 % ee (R), ZnL4 and ZnL5 afforded good chemical yields and high enantioselectivities of up to 89 % ee (R). UV titrations of proline and ZnL4 with acetylacetone (acac) in DMSO/H2O (1:2) indicate that ZnL4 facilitates the formation of the ZnL4 ? (acac)? complex (Kapp=2.1×102 M ?1), whereas L ‐proline forms a Schiff base with acac with a very small equilibrium constant. These results suggest that the amino acid components and the Zn2+ ions in ZnL4 and ZnL5 function in a cooperative manner to generate the Zn2+‐enolate of acetone, thus permitting efficient enantioselective C? C bond formation with aldehydes.  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