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 93 毫秒
1.
A 1:1 complex of mercuric chloride with D-peniccillamine has been isolated and characterised as 2[(μ3-Cl){HgSC(CH3)2CH(NH3)COO}3]·3(μ2-Cl)·2(H3O)·(H2O·Cl)3. The compound crystallises in cubic space group P4132, with a = 18.679(5) Å and Z = 4. The structure, refined to RF = 0.086 for 443 observed Mo-Kα diffractometer data, features a triply bridging chloride ion linking three equivalent [HgSC(CH3)2CH(NH3)COO]+ units [Hg-Cl = 2.37(1) Å, Hg-Cl-Hg′ = 98.5(9)°]. The carboxylate groups of a pair of adjacent penicillamine ligands are strongly linked via a symmetrical O?H?O hydrogen bond of length 2.24(8) Å, and neighboring pyramidal trinuclear [μ3-Cl){HgSC(CH3)2CH(NH3)-COO}3]2+ moieties are further connected by symmetrical chloride bridges [Hg-Cl = 3.06(2) Å; HgClHg′' = 79.6(7)°] to form a three-dimensional network. The voids in the lattice are filled by hydronium ions and novel planar cyclic hydrogen-bonded (H2O·Cl?)3 rings of edge O-H?Cl = 2.46(4) Å.  相似文献   

2.
Chiral ionic liquids (CILs) with amino acids as cations have been applied as novel chiral ligands coordinated with Cu2+ to separate tryptophan enantiomers in ligand exchange chromatography. Four kinds of amino acid ionic liquids, including [L‐Pro][CF3COO], [L‐Pro][NO3], [L‐Pro]2[SO4], and [L‐Phe][CF3COO] were successfully synthesized and used for separation of tryptophan enantiomers. To optimize the separation conditions, [L‐Pro][CF3COO] was selected as the model ligand. Some factors influencing the efficiency of chiral separation, such as copper ion concentration, CILs concentration, methanol ratio (methanol/H2O, v/v), and pH, were investigated. The obtained optimal separation conditions were as follows: 8.0 mmol/L Cu(OAc)2, 4.0 mmol/L [L‐Pro][CF3COO] ,and 20% (v/v) methanol at pH 3.6. Under the optimum conditions, acceptable enantioseparation of tryptophan enantiomers could be observed with a resolution of 1.89. The results demonstrate the good applicability of CILs with amino acids as cations for chiral separation. Furthermore, a comparative study was also conducted for exploring the mechanism of the CILs as new ligands in ligand exchange chromatography. Chirality 26:160–165, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
Fluorotelomer alcohols [FTOHs, F(CF2)nCH2CH2OH, n = 4, 6, and 8] are emerging environmental contaminants. Biotransformation of FTOHs by mixed bacterial cultures has been reported; however, little is known about the microorganisms responsible for the biotransformation. Here we reported biotransformation of FTOHs by two well‐studied Pseudomonas strains: Pseudomonas butanovora (butane oxidizer) and Pseudomonas oleovorans (octane oxidizer). Both strains could defluorinate 4:2, 6:2, and 8:2 FTOHs, with a higher degree of defluorination for 4:2 FTOH. According to the identified metabolites, P. oleovorans transformed FTOHs via two pathways I and II. The pathway I led to the production of x:2 ketone [dominant metabolite, F(CF2)xC(O)CH3; x = n ? 1, n = 6 or 8], x:2 sFTOH [F(CF2)xCH(OH)CH3], and perfluorinated carboxylic acids (PFCAs, perfluorohexanoic, or perfluorooctanoic acid). The pathway II resulted in the formation of x:3 polyfluorinated acid [F(CF2)xCH2CH2COOH] and relatively minor shorter‐chain PFCAs (perfluorobutyric or perfluorohexanoic acid). Conversely, P. butanovora transformed FTOHs by using the pathway I, leading to the production of x:2 ketone, x:2 sFTOH, and PFCAs. This is the first study to show that individual bacterium can bio‐transform FTOHs via different or preferred transformation pathways to remove multiple ? CF2? groups from FTOHs to form shorter‐chain PFCAs. Biotechnol. Bioeng. 2012; 109: 3041–3048. © 2012 Wiley Periodicals, Inc.  相似文献   

4.
The macrocyclisation reaction of 3,3′-(3,6-dioxaoctane-1,8-diyldioxy)-bis(2-hydroxybenzaldehyde) (1) with S-methylisothiosemicarbazide hydroiodide (H2NNC(SCH3)NH2·HI) in the presence of potassium triflate, followed by addition of M(CH3COO)2·nH2O, where M=Ni, Cu, Zn, afforded [NiLKI3] (2), [NiLK(CF3SO3)] (3), [CuLK(CF3SO3)(CH3OH)] (4) and [(ZnILK)2CH3OH] (5), respectively. Compounds 2-5 have been characterised by X-ray crystallography. IR, electronic, mass, 1H, 13C{1H} and 19F{1H} NMR spectra are reported. Magnetic susceptibility measurements and ESR spectra of 4 indicate weak intermolecular spin-spin interactions, which are mostly dipolar in origin.  相似文献   

5.
《Inorganica chimica acta》2006,359(8):2448-2454
The three novel gadolinium(III) containing compounds (NH3CH3)[Gd(CF3CF2COO)4(H2O)] (1), (NH3C2H5)[Gd(CF3CF2COO)4(H2O)] (2) and ((CH3)4N)[Gd(CF3CF2COO)3(H2O)2]CF3CF2COO (3) were synthesized and structurally characterized by X-ray crystallography. In the crystal structures of 1 and 2, the gadolinium ions are bridged by carboxylate groups to dimers with a Gd3+–Gd3+ distance of 451.6(2) (1) and 451.8(3) pm (2), respectively. In the crystal structure of 3 the Gd3+ ions are bridged by carboxylate groups to chains with almost the same Gd3+–Gd3+ distances (494.0(8) and 503.4(7) pm). The magnetic behaviour of 1 and 2 was investigated in the temperature range of 1.76–300 K. The magnetic data indicate weak antiferromagnetic interactions within the dimeric unit.  相似文献   

6.
The dimer [Ir(μ-Cl)(C8H14)2]2 reacts with the ligands (S)-(C5H4CH2CH(Ph)PPh2)Li and (R)-(C5H4CH(Cy)CH2PPh2)Li to give (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(C8H14)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(C8H14)], which upon treatment with CH3I at room temperature afford the cationic iridium(III) compounds (S,SIr)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(C8H14)][I] as a single diastereomer, and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(C8H14)][I] as a 9:1 mixture of two diastereomers. If the oxidative addition reaction is performed at reflux in methylene chloride, the starting complexes convert to the neutral compounds (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(I)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(I)] as 1.6:1 and 3.3:1 mixtures of diastereoisomers, respectively. Carbonyl iridium complexes are synthesized by reacting [IrCl(CO)(PPh3)2] with the ligands to afford (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CO)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CO)]. They give upon treatment with CH3I the cationic species (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(CO)][I] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(CO)][I] as 1.6:1 and 3:1 mixture of diastereomers, respectively. No migratory-insertion of the methyl group into the carbonyl-metal bond has been observed even after prolonged heating.  相似文献   

7.
The chiral recognition property of poly[(1→6)-2,5-anhydro-3,4-di-O-alkyl-D-glucitol] ( 1 ) toward racemic RCH (CO2CH3)NH3+ · PF6? ( 2 · HPF6) has been studied using a transport system involving an aqueous source and receiving phases separated by a chloroform phase containing 1 . Transport rates for aromatic guests 2a (R = Ph) and 2b (R = CH2Ph) were faster than those for aliphatic guests, 2c (R = CH(CH3)2) and 2d (R = CH2CH(CH3)2), using the polymer substituted with methyl groups ( 1a ). The enantiomeric excess (e.e.) was 10.9% for 2a as a maximum value and decreased in the order of 2a > 2c > 2b = 2d . When the transport of 2a · HPF6 was carried out using the polymers with 3,4-di-O-methyl ( 1a ), ethyl ( 1b ), allyl ( 1c ), and pentyl ( 1d ) groups, the e.e. was 22.0% for 1d as a maximum value and increased in the order of 1a < 1b < 1c < 1d . The formation of a complex between 1a and 2a · HPF6 was confirmed by 1H and 13C NMR spectral measurements. © 1995 Wiley-Liss, Inc.  相似文献   

8.
A series of 30 RCO–HfR–NH2 derivatives show preference for the mouse MC1R vs MC3-5Rs. trans-4-HOC6H4CHCHCO–HfR–NH2 (13) [EC50 (nM): MC1R 83, MC3R 20500, MC4R 18130 and MC5R 935; ratio 1:246:217:11] is 11 times more potent than the lead compound LK-394 Ph(CH2)3CO–HfR–NH2 (2) and only 11 times less potent than the native tridecapeptide α-MSH at mMC1R. Differences in conformations of 2 and 13 are discussed.  相似文献   

9.
《Inorganica chimica acta》1988,149(2):259-264
The bis(N-alkylsalicylaldiminato)nickel(II) complexes Ni(R-sal)2 with R = CH(CH2OH)CH(OH)Ph (I), R = CH(CH3)CH(OH)Ph (II) and R = CH2CH2Ph (III; Ph = phenyl) were prepared and characterized. In the solid state I and II are paramagnetic (μ = 3.2 and 3.3 BM at 20 °C, respectively), whereas III is diamagnetic. It follows from the UV-Vis spectra that in acetone solution I is six-coordinate octahedral and III is four-coordinate planar, the spectrum of II showing characteristics of both modes of coordination. Vis spectrophotometry and stopped-flow spectrophotometry were applied to study the kinetics of ligand substitution in I–III by H2salen (= N,N′-disalicylidene-ethylenediamine) in the solvent acetone at different temperatures. The kinetics follow a second-order rate law, rate = k[H2-salen] [complex]. At 20 °C the sequence of rate constants is k(III):k(II):k(I) = 11 850:40.6:1. The activation parameters are ΔH(I) = 112, ΔH(II) = 40.7, ΔH(III) = 35.7 kJ mol−1 and ΔS(I) = 92, ΔS(II) = −103, ΔS(III) = −89 J K−1 mol−1. The enormous difference in rate between complexes I, II and III, which is less pronounced in methanol, is attributed to the existence of a fast equilibrium planar ⇌ octahedral, which is established in the case of I and II by intramolecular octahedral coordination through the hydroxyl groups present in the organic group R. An A-mechanism is suggested to control the substitution in the sense that the entering ligand attacks the four-coordinate planar complex, the octahedral complex being kinetically inert.  相似文献   

10.
A house fly attracting substance, referred to as D3 in the preceeding paper,1) was identified with 1,3-diolein.

Among the related compounds, 1- and 2-monoolein and α,ω-glycol monooleate with the formula: CH3(CH2)7CH=CH(CH2)7COO(CH2)nOH (n≦6), were found to have activities ten times that of 1,3-diolein.  相似文献   

11.
When aqueous solutions containing [Ru(NH2CH3)6]2+ are exposed to oxygen [Ru(NH2CH3)5(OH]2+ is produced as an identifiable intermediate as a result of the slow replacement of co-ordinated methylamine by water, and the subsequent rapid oxidation of this ruthenium(II) aquo complex. The [Ru(NH2CH3)5(OH)]2+ ion then undergoes another slow reaction, probably the replacement of another methylamine ligand by water. All subsequent reactions leading to Ru(CN)3·3H2O are rapid. Rate data are reported, and a mechanism involving β-elimination from co-ordinated methylamine is postulated.  相似文献   

12.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

13.
Tris(triazolyl)borate ligands (Ttz) of intermediate steric bulk were synthesized to investigate their potential for hydrogen bonding and improved solubility in hydrophilic solvents as applied to biomimetic chemistry. The crystal structure of 3-phenyl-5-methyl-1,2,4-triazole (HtzPh,Me) revealed hydrogen bonding and π stacking interactions. The new ligand salt, potassium tris(3-phenyl-5-methyl-1,2,4-triazolyl)borate (KTtzPh,Me) was synthesized as the first example of a Ttz ligand of intermediate steric bulk. Metathesis between KTtzPh,Me and NaCl followed by recrystallization produced [NaTtzPh,Me] · 6CH3OH in which the geometry around the sodium is octahedral with an unusual N3O3 donor set; this structure also shows that a hydrogen bonding network is formed by methanol molecules and triazole nitrogens. (TtzPh,Me)ZnCl was synthesized and characterized crystallographically as [(TtzPh,Me)ZnCl] · 0.5CH3OH in which the zinc is tetrahedral and the triazole rings are within hydrogen bonding distance of CH3OH. All of these new compounds are methanol soluble to varying degrees and HtzPh,Me and KTtzPh,Me are soluble in methanol/water mixtures.  相似文献   

14.
Three silver(I) complexes of dibenzo-18-crown-6-ether (DB[18]C6), [Ag(DB[18]C6)(ClO4)](THF) (1), [Ag(DB[18]6)(CF3SO3)]2(acetone)2 (2) and [Ag(DB[18]C6)(CF3COO)]2(AgCF3COO)2 (3) have been synthesized in different solvents and characterized structurally. In each complex, silver ions prefer an octahedral coordination geometry and form close dinuclear complex with DB[18]C6 based on cation-π interaction in η2-fashion. In particular, the coordination unit involving σ bonding at an oxygen group and π-π bonding between two benzene rings is quite unique.  相似文献   

15.
Ferrocene reacts with hexafluoroacetone trihydrate in refluxing octane to afford >80% yields of [CpFe(η5-C5H4C(CF3)2OH)] (X-ray), carrying out the reactions at 180 °C gives an additional 5% yield of [Fe(η5-C5H4C(CF3)2OH)2] (X-ray).The mono alcohol is lithiated with ButOK/BunLi/TMEDA affording partial conversion to mixtures of [CpFe(1,2-η5-C5H3C(CF3)2OH)(X)] and [Fe(η5-C5H4X)(1,2-η5-C5H3C(CF3)2OH)(X)] (X = SMe, CPh2OH) upon reaction with Me2S2 or OCPh2.For X = CPh2OH both structures are crystallographically characterised.Enantiopure [CpFe(1,2-η5-C5H3C(CF3)2OH)(SMe)] can be prepared from (R)-[CpFe(η5-C5H4S(O)C6H4Me)] via [CpFe(1,2-η5-C5H3S(O)C6H4Me)(C(CF3)2OH)] (X-ray) or [CpFe(1,2-η5-C5H3S(O)C6H4Me)(SMe)].Related procedures allow the preparation of [CpFe(1,2-η5-C5H3CPh2OH)(Y)] (Y = SMe, CHO (X-ray), C(CF3)2OH) and[CpFe(1,2-η5-C5H3C(CF3)2OH)(CHO)].  相似文献   

16.
《Inorganica chimica acta》1988,149(2):253-258
The chiroptical properties of five-coordinate diastereomeric complexes of general formula [PtCl2(R,R)-{C6H5CH(CH3)N(CH3)CH2}2{olefin}], with olefin ligands having electron-withdrawing substituents, have been investigated. The sign of CD bands in the 28 000–30 000 cm−1 region appears to be correlated to the absolute configuration of the prochiral coordinated alkene. Single-crystal X-ray diffraction structure determination has been performed on the single diastereomer [PtCl2(E-but-2-enedinitrile)(R,R)-{C6H5CH(CH3)N(CH3)CH2}2]· C6H6. The compound crystallizes in the monoclinic space group C2 with a = 17.842(2), b = 8.466(1), c = 10.464(1) Å, β = 109.34(1)°, Z = 2. The number of observed reflections was 1943 and the final R and Rw values were 0.020 and 0.028 respectively. Trigonal-bipyramidal geometry is observed around the Pt atom, with the two Cl atoms in axial positions. The unsaturated ligand lies in the equatorial plane disclosing S,S absolute configuration.  相似文献   

17.
Magnesium dimers play important roles in inorganic and organometallic chemistry. This study evaluates the inherent bridging ability of a range of different ligands in magnesium dimers. In the first part, the Cambridge Structural Database is interrogated to establish the frequency of different types of ligands found in bridging versus terminal positions in two key structural motifs: one in which there are two bridging ligands (the D 2h “Mg2(μ-X2)” structure); the other in which there are three bridging ligands (the C 3v “Mg2(μ-X3)” structure). The most striking finding from the database search is the overwhelming preference for magnesium dimers possessing two bridging ligands. The most common bridging ligands are C-, N-, and O-based. In the second part, DFT calculations (at the B3LYP/6-311+G(d) level of theory) are carried out to examine a wider range of structural types for dimers consisting of the stoichiometries Mg2Cl3R and Mg2Cl2R2, where R = CH3, SiH3, NH2, PH2, OH, SH, CH2CH3, CH=CH2, C≡CH, Ph, OAc, F and Br. Consistent with the database search, the most stable magnesium dimers are those that contain two bridging ligands. Furthermore, it was demonstrated that the electronic effect of the bridging ligands is important in influencing the stability of the magnesium dimers. The preference for a bridging ligand, which reflects its ability to stabilize a magnesium dimer, follows the order: OH > NH2 > C≡CH > SH > Ph > Br > PH2 = CH=CH2 > CH2CH3 > CH3 > SiH3. Finally, the role that the ether solvent Me2O has on the stability of isomeric Mg2Cl2Me2 dimers was studied. It was found that the first solvent molecule stabilizes the dimers, while the second solvent molecule can either have a stabilizing or destabilizing effect, depending on the isomer structure.  相似文献   

18.
Two new salts based on heterocyclic organic cations and uranyl triacetate anion were obtained via reaction of zinc uranyl acetate with 2-substituted imidazoles in presence of an excess of acetic acid. Uranyl triacetate anion in [2-MeImH]+ [UO2(CH3COO)3] and [2-PhImH]+ [UO2(CH3COO)3] H2O has an expected bipyramidal structure with linear uranyl group and three acetate groups laying in equatorial plane. [2-MeImH]+ [UO2(CH3COO)3] structure analysis reveals H-bonded 1D chains connected through N-H···O hydrogen bonds. 2-phenylimidazolium in [2-PhImH]+ [UO2(CH3COO)3] H2O demonstrate planar geometry without any rotation of its rings, which was not registered before. H-bonds and π-π interactions of phenyl groups in this system lead to complicate 2D “sandwich” layer formation. The main features of IR- and luminescence spectrum of both compounds are also discussed.  相似文献   

19.
The structures of MoO2[NH2C(CH3)2CH2S]2 and MoO2[SC(CH3)2CH2NHCH2CH2NHCH2C(CH3)2S] have been determined using X-ray diffraction intensity data collected by counter techniques. MoO2[NH2C(CH3)2CH2S]2 crystallizes in space group Pbca with a = 11.234(3), b = 11.822(3) and c = 20.179(5) Å, V = 2680(2) Å3 and Z = 8. Its structure is derived from octahedral coordination with cis oxo groups [MoO = 1.705(3) and 1.705(3)], trans thiolate donors cis to the oxo groups [MoS = 2.416(1) and 2.402(1) and N donors trans to oxo [MoN = 2.325(3) and 2.385(4) Å]. MoO2[SC(CH3)2CH2NHCH2CH2NHCH2C(CH3)2S] crystallizes in the space group P21/c with a = 10.798(5), b = 6.911(2), c = 20.333(9) Å, β = 95.20°, V = 1511(2) Å3 and Z = 4. Its structure is very similar to that of MoO2[NH2C(CH3)2CH2S]2 with MoO = 1.714(2) and 1.710(2), MoS = 2.415(1) and 2.404(1) and MoN = 2.316(3) and 2.362(3). The small differences in the geometries of the two compounds are attributed to the constraints of the extra chelate ring in the complex with the tetradentate ligand. The structures in this paper stand in contrast to those reported for complexes of similar ligands wherein steric hindrance produces complexes with a skew trapezoidal bipyramidal structure.  相似文献   

20.
Factors have been investigated which govern the electrophilic reactivity of alkyl halides with thiolate anions in aqueous solution. In the series of alkyl halides studied, some are potential metal-directed affinity labels, while others are frequently used in protein modification. Previous data on the kinetics of this type of alkylation are compared with the present results. The influence of electronic, polar, and steric factors on alkyl halide reactivity is seen. The following order of reactivity for alkyl halides bearing different α substituents was observed: RCH2CH(X)COOCH3 > RCH2CH(X)CONH2 > RCH2CH(X)COOH > RCH2CH2X > RCH2CH(X)CH2OH. The metal-directed affinity labels are imidazole derivatives, some of which have substituents in their imidazole ring. The effect of the imidazole ring and of ring substitution on reactivity is seen. The nucleophilic reactivity of thiols is highly pH dependent since the thiolate anion (RS?) is the reactive species, but only minor differences emerged between different free thiolates.  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