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 46 毫秒
1.
Three new ligands and their palladium(II) complexes of general formula [PdCl2(R2-S,S-eddp)] (R = n-propyl, n-butyl and n-pentyl) have been synthesized and characterized by microanalysis, infrared and 1H and 13C NMR spectroscopy. Antimicrobial activity of these ligands and complexes was tested by microdilution method and both minimal inhibitory and microbicidal concentration were determined. These tested complexes demonstrated the significant antifungal activity against pathogenic fungi Aspergillus flavus and Aspergillus fumigatus. On the other hand, these complexes demonstrated moderate antibacterial activity.  相似文献   

2.
Two isomers (R,S,R,S- and R,R,S,S-) of five coordinate complex [Cu(L)Cl]+ have been separated and characterised. These two isomers have significantly different spectrochemical and electrochemical properties. Absorption maximum of R,S,R,S-[Cu(L)Cl]+ shifts to longer wavelength and its reduction potential shifts to more positive direction comparing those of R,R,S,S-[Cu(L)Cl]+. R,S,R,S-[Cu(L)Cl]+ is significantly distorted to trigonal-bipyramidal structure, whereas R,R,S,S-[Cu(L)Cl]+ retains almost square-planar geometry. The average bond distance of Cu-N in basal plane of R,S,R,S-[Cu(L)Cl]+ is longer by 0.024 Å than that of R,R,S,S-[Cu(L)Cl]+, whereas the bond distance of Cu-Cl in former is shorter by 0.200 Å than that in latter. The isolated square-planar complexes of R,R,S,S- and R,S,R,S-[Cu(L)](ClO4)2 are converted to the R,R,S,S- and R,S,R,S-[Cu(L)Cl]+ by the addition of Cl in nitromethane solution with the rate constants, k=1.70 (±0.02) and 8.31 (±0.07) M−1 s−1, respectively.  相似文献   

3.
Thiourea, PhNHC(S)NHP(O)(OPri)2 (LH) chelates of CoII, NiII, and PdII ions have been obtained and investigated by single-crystal X-ray diffraction, UV, IR, NMR spectroscopy, and EI mass-spectrometry. The unusual 1,3-N,S-coordination via sulfur and NP(O) nitrogen atoms has been found in the trans-square-planar NiL2 and PdL2 complexes, whereas the 1,5-O,S-coordination is realized in the tetrahedral CoL2 complex. DFT calculations have revealed significant stabilization of the 1,3-N,S-structures due to stronger crystal field and the NH-OP hydrogen bonds.  相似文献   

4.
Huang HL  Xu YJ  Liu HL  Liu XQ  Shang JN  Han GT  Yao MJ  Yuan CS 《Phytochemistry》2011,72(6):514-517
A dimeric eremophilane sesquiterpene lactone with a cyclobutane ring, biliguhodgsonolide (1) and an uncommon seco-sesquiterpene derivative, (4S,5S,6R,10R)-8,9-seco-12-hydroxyeremophil-7(11)-en-14,6;12,8-diolid-9-al (2), were isolated from the roots and rhizomes of Ligularia hodgsonii Hook. Their structures, including the absolute stereochemistry, were elucidated by spectroscopic data and CD analysis. The cyclobutane ring was confirmed by single-crystal X-ray diffraction.  相似文献   

5.
New tetradentate ligands 2-(2-mercaptoethylthio)-N-(pyridin-2-ylmethyl)acetamide H2L1 and 2-chloro-2-(2-mercaptoethylthio)-N-(pyridin-2-ylmethyl)acetamide H2L2 were synthesised from the reaction of 2-aminomethanepyridine with 1,4-dithian-2-one and 3-chloro-1,4-dithian-2-one, respectively. Monomeric complexes of these ligands, of general formulae K[CrIII(Ln)Cl2], K2[MnII(Ln)Cl2] and [M(Ln)] (M = Fe(II), Co(II), Ni(II), Cu(II), Zn(II), Cd(II) or Hg(II); n = 1, 2) are reported. The mode of bonding and overall geometry of the complexes were determined through IR, UV-Vis, NMR and mass spectral studies, magnetic moment measurements, elemental analysis, metal content and conductance. These studies revealed octahedral geometries for the Cr(III), Mn(II) complexes, square planar for Ni(II) and Cu(II) complexes and tetrahedral for the Fe(II), Co(II), Zn(II), Cd(II) and Hg(II) complexes. The study of complex formation via molar ratio in DMF solution has been investigated and results were consistent to those found in the solid complexes with a ratio of (M:L) as (1:1).  相似文献   

6.
The inclusion of (R)- and (S)-camphor compounds in α-cyclodextrin has been studied by X-ray crystallography. The crystal structures of the complexes reveal that one guest molecule is accommodated inside the cavity formed by a head-to-head cyclodextrin dimer. In the crystal lattice, the dimers form layers which are successively shifted by half a dimer. In both (R)- and (S)-cases, the camphor molecule exhibits disorder and occupies three major sites with orientations that can be described as either ‘polar’ or ‘equatorial’. Molecular dynamics simulations performed for the observed complexes indicate that although the carbonyl oxygen of both (R)- and (S)-camphor switches between different hydrogen bonding partners, it maintains the observed mode of ‘polar’ or ‘equatorial’ alignment.  相似文献   

7.
8.
Ketol-acid reductoisomerase (KARI; EC 1.1.1.86) is an enzyme in the branched-chain amino acid biosynthesis pathway where it catalyzes the conversion of 2-acetolactate into (2R)-2,3-dihydroxy-3-isovalerate or the conversion of 2-aceto-2-hydroxybutyrate into (2R,3R)-2,3-dihydroxy-3-methylvalerate. KARI catalyzes two reactions—alkyl migration and reduction—and requires Mg2+ and NADPH for activity. To date, the only reported structures for a plant KARI are those of the spinach enzyme-Mn2+-(phospho)ADP ribose-(2R,3R)-2,3-dihydroxy-3-methylvalerate complex and the spinach KARI-Mg2+-NADPH-N-hydroxy-N-isopropyloxamate complex, where N-hydroxy-N-isopropyloxamate is a predicted transition-state analog. These studies demonstrated that the enzyme consists of two domains, N-domain and C-domain, with the active site at the interface of these domains. Here, we have determined the structures of the rice KARI-Mg2+ and rice KARI-Mg2+-NADPH complexes to 1.55 Å and 2.80 Å resolutions, respectively. In comparing the structures of all the complexes, several differences are observed. Firstly, the N-domain is rotated up to 15° relative to the C-domain, expanding the active site by up to 4 Å. Secondly, an α-helix in the C-domain that includes residues V510-T519 and forms part of the active site moves by ∼ 3.9 Å upon binding of NADPH. Thirdly, the 15 C-terminal amino acid residues in the rice KARI-Mg2+ complex are disordered. In the rice KARI-Mg2+-NADPH complex and the spinach KARI structures, many of the 15 residues bind to NADPH and the N-domain and cover the active site. Fourthly, the location of the metal ions within the active site can vary by up to 2.7 Å. The new structures allow us to propose that an induced-fit mechanism operates to (i) allow substrate to enter the active site, (ii) close over the active site during catalysis, and (iii) open the active site to facilitate product release.  相似文献   

9.
The repellent activity of the essential oil of the catmint plant, Nepeta cataria (Lamiaceae), and the main iridoid compounds (4aS,7S,7aR) and (4aS,7S,7aS)-nepetalactone, was assessed against (i) major Afro-tropical pathogen vector mosquitoes, i.e. the malaria mosquito, Anopheles gambiae s.s. and the Southern house mosquito, Culex quinquefasciatus, using a World Health Organisation (WHO)-approved topical application bioassay (ii) the brown ear tick, Rhipicephalus appendiculatus, using a climbing repellency assay, and (iii) the red poultry mite, Dermanyssus gallinae, using field trapping experiments. Gas chromatography (GC) and coupled GC-mass spectrometry (GC-MS) analysis of two N. cataria chemotypes (A and B) used in the repellency assays showed that (4aS,7S,7aR) and (4aS,7S,7aS)-nepetalactone were present in different proportions, with one of the oils (from chemotype A) being dominated by the (4aS,7S,7aR) isomer (91.95% by GC), and the other oil (from chemotype B) containing the two (4aS,7S,7aR) and (4aS,7S,7aS) isomers in 16.98% and 69.83% (by GC), respectively. The sesquiterpene hydrocarbon (E)-(1R,9S)-caryophyllene was identified as the only other major component in the oils (8.05% and 13.19% by GC, respectively). Using the topical application bioassay, the oils showed high repellent activity (chemotype A RD50 = 0.081 mg cm−2 and chemotype B RD50 = 0.091 mg cm−2) for An. gambiae comparable with the synthetic repellent DEET (RD50 = 0.12 mg cm−2), whilst for Cx. quinquefasciatus, lower repellent activity was recorded (chemotype A RD50 = 0.34 mg cm−2 and chemotype B RD50 = 0.074 mg cm−2). Further repellency testing against An. gambiae using the purified (4aS,7S,7aR) and (4aS,7S,7aS)-nepetalactone isomers revealed overall lower repellent activity, compared to the chemotype A and B oils. Testing of binary mixtures of the (4aS,7S,7aR) and (4aS,7S,7aS) isomers across a range of ratios, but all at the same overall dose (0.1 mg), revealed not only a synergistic effect between the two, but also a surprising ratio-dependent effect, with lower activity for the pure isomers and equivalent or near-equivalent mixtures, but higher activity for non-equivalent ratios. Furthermore, a binary mixture of (4aS,7S,7aR) and (4aS,7S,7aS) isomers, in a ratio equivalent to that found in chemotype B oil, was less repellent than the oil itself, when tested at two doses equivalent to 0.1 and 0.01 mg chemotype B oil. The three-component blend including (E)-(1R,9S)-caryophyllene at the level found in chemotype B oil had the same activity as chemotype B oil. In a tick climbing repellency assay using R. appendiculatus, the oils showed high repellent activity comparable with data for other repellent essential oils (chemotype A RD50 = 0.005 mg and chemotype B RD50 = 0.0012 mg). In field trapping assays with D. gallinae, addition of the chemotype A and B oils, and a combination of the two, to traps pre-conditioned with D. gallinae, all resulted in a significant reduction of D. gallinae trap capture. In summary, these data suggest that although the nepetalactone isomers have the potential to be used in human and livestock protection against major pathogen vectors, intact, i.e. unfractionated, Nepeta spp. oils offer potentially greater protection, due to the presence of both nepetalactone isomers and other components such as (E)-(1R,9S)-caryophyllene.  相似文献   

10.
Dinuclear tris(acetylacetonato)ruthenium(III) complexes bridged by one sulfur atom (1) or two sulfur atoms (2) at the γ-position of the acetylacetonate have been synthesized by the reactions of tris(acetylacetonato)ruthenium(III) with SCl2 or S2Cl2. The molecular structure of 2 has been determined by single crystal X-ray diffraction study. The cyclic voltammograms of both the dinuclear complexes exhibit two one-electron reduction waves in acetonitrile (AN), dichloromethane (DM), benzonitrile (BN), and N,N-dimethylformamide (DMF). While the complex 1 exhibited two one-electron oxidation waves in AN, DM, and BN, complex 2 showed only irreversible waves in all the solvents. The comproportionation constants (Kc) for mixed-valence states of both RuII/RuIII and RuIII/RuIV were calculated from the redox potentials of dinuclear complexes. The values of log10Kc (RuII/RuIII) (for complexes 1 and 2) and log10Kc (RuIII/RuIV) (for complex 1) were between 1.35 and 3.55. These values are not so large and hence, the complexes may be classified as class II in the Robin and Day classification. Although no relationship could be found between Kc and the dielectric constant of the solvent, there exists a correlation between donor number (DN) of the solvent and log10Kc values.  相似文献   

11.
trans-Dioxoruthenium(VI) porphyrin complexes have been developed as one of the best-characterized model systems for heme-containing enzymes. Traditionally, this type of compounds can be prepared by oxidation of ruthenium(II) precursors with peroxyacids and other terminal oxidants under different conditions, depending on the porphyrin ligands. In this work, a new photochemical generation of trans-dioxoruthenium(VI) porphyrins has been developed by extension of the known photo-induced ligand cleavage reactions. Refluxing ruthenium(II) carbonyl porphyrins [RuII(Por)(CO)] in carbon tetrachloride afforded dichlororuthenium(IV) complexes [RuIV(Por)Cl2]. Facile exchange of the counterions in [RuIV(Por)Cl2] with Ag(ClO3) or Ag(BrO3) gave the corresponding dichlorate [RuIV(Por)(ClO3)2] or dibromate [RuIV(Por)(BrO3)2] salts. Visible-light photolysis of the photo-labile porphyrin-ruthenium(IV) dichlorates or dibromates resulted in homolytic cleavage of the two O-Cl or O-Br bonds in the axial ligands to produce trans-dioxoruthenium(IV) species [RuVI(Por)O2] bearing different porphyrin ligands.  相似文献   

12.
The reactions of [PtMe3(OAc)(bpy)] (4) with the N,S and S,S containing heterocycles, pyrimidine-2-thione (pymtH), pyridine-2-thione (pytH), thiazoline-2-thione (tztH) and thiophene-2-thiol (tptH), resulted in the formation of the monomeric complexes [PtMe3(-κS)(bpy)] ( = pymt, 5; pyt, 6; tzt, 7; tpt, 8), where the heterocyclic ligand is coordinated via the exocyclic sulfur atom. In contrast, in the reactions of [PtMe3(OAc)(Me2CO)x] (3, x = 1 or 2) with pymtH, pytH, tztH and tptH dimeric complexes [{PtMe3(μ-)}2] (μ- = pymt, 9; pyt, 10; tzt, 11) and the tetrameric complex [{PtMe33-tpt-κS)}4] (12), respectively, were formed. The complexes were characterized by microanalyses, 1H and 13C NMR spectroscopy and negative ESI-MS (12) measurements. Single-crystal X-ray diffraction analysis of [PtMe3(pymt-κS)(bpy)] (5) exhibited a conformation where the pymt ligand lies nearly perpendicular to the complex plane above the bpy ligand that was also confirmed by quantum chemical calculations on the DFT level of theory.  相似文献   

13.
5′-R and 5′-S diastereoisomers of 8,5′-cyclo-2′-deoxyadenosine (cdA) and 8,5′-cyclo-2′-deoxyguanosine (cdG) containing a base-sugar covalent bond are formed by hydroxyl radicals. R-cdA and S-cdA are repaired by nucleotide excision repair (NER) in mammalian cellular extracts. Here, we have examined seven purified base excision repair enzymes for their ability to repair S-cdG or S-cdA. We could not detect either excision or binding of these enzymes on duplex oligonucleotide substrates containing these lesions. However, both lesions were repaired by HeLa cell extracts. Dual incisions by human NER on a 136-mer duplex generated 24–32 bp fragments. The time course of dual incisions were measured in comparison to cis-anti-B[a]P-N2-dG, an excellent substrate for human NER, which showed that cis-anti-B[a]P-N2-dG was repaired more efficiently than S-cdG, which, in turn, was repaired more efficiently than S-cdA. When NER efficiency of S-cdG with different complementary bases was investigated, the wobble pair S-cdG·dT was excised more efficiently than the S-cdG·dC pair that maintains nearly normal Watson-Crick base pairing. But S-cdG·dA mispair with no hydrogen bonds was excised less efficiently than the S-cdG·dC pair. Similar pattern was noted for S-cdA. The S-cdA·dC mispair was excised much more efficiently than the S-cdA·dT pair, whereas the S-cdA·dA pair was excised less efficiently. This result adds to complexity of human NER, which discriminates the damaged base pairs on the basis of multiple criteria.  相似文献   

14.
Reaction of copper(II) acetate with the (S)-enantiomer of a tridentate binaphthyl Schiff base ligand, 2-(3,5-dichloro-2-hydroxybenzylideneamino)-2′-hydroxy-1,1′-binaphthyl (H2L), in methanol afforded mononuclear copper(II) complex [CuII(HL)2] ((S,S)-1) in 52% isolated yield. The same reaction gave dinuclear copper(II) complex [CuII2(L)2] ((R,S)-2) in 73% isolated yield when racemic-H2L was used instead of (S)-H2L. Both complexes (S,S)-1 and (R,S)-2 were characterized by elemental analysis, mass spectrometry, and X-ray crystallography. The present work highlights the functioning of ligand chirality as a ‘switch’ for selective formation of mono- and dinuclear metal complexes.  相似文献   

15.
The cytosolic protein CERT transfers ceramide from the endoplasmic reticulum to the Golgi apparatus where ceramide is converted to SM. The C-terminal START (steroidogenic acute regulatory protein-related lipid transfer) domain of CERT binds one ceramide molecule in its central amphiphilic cavity. (1R,3R)-N-(3-Hydroxy-1-hydroxymethyl-3-phenylpropyl)alkanamide (HPA), a synthesized analogue of ceramide, inhibits ceramide transfer by CERT. Here we report crystal structures of the CERT START domain in complex with HPAs of varying acyl chain lengths. In these structures, one HPA molecule is buried in the amphiphilic cavity where the amide and hydroxyl groups of HPA form a hydrogen-bond network with specific amino acid residues. The Ω1 loop, which has been suggested to function as a gate of the cavity, adopts a different conformation when bound to HPA than when bound to ceramide. In the Ω1 loop region, Trp473 shows the largest difference between these two structures. This residue exists inside of the cavity in HPA-bound structures, while it is exposed to the outside of the protein in the apo-form and ceramide-bound complex structures. Surface plasmon resonance experiments confirmed that Trp473 is important for interaction with membranes. These results provide insights into not only the molecular mechanism of inhibition by HPAs but also possible mechanisms by which CERT interacts with ceramide.  相似文献   

16.
Pentaammineruthenium(III) complexes of deoxyinosine (dIno) and xanthosine (Xao) ([RuIII(NH3)5(L)], L?is?dIno, Xao) in basic solution were studied by UV?Cvis spectroscopy, liquid chromatography/electrospray ionization mass spectrometry, and high-performance liquid chromatography. Both RuIII complexes disproportionate to RuII and RuIV. Disproportionation followed the rate law d[RuII]/dt?=?(k o?+?k 1[OH?])[RuIII]. k o and k 1 of disproportionation at 25?°C were 2.1 (±0.1)?×?10?3?s?1 and 21.4?±?3.2?M?1 s?1, respectively, for [RuIII(NH3)5(dIno)], and 3.5 (±0.7)?×?10?4?s?1 and 59.7?±?3.6?M?1?s?1, respectively, for [RuIII(NH3)5(Xao)]. The [RuIII(NH3)5(Xao)] complex disproportionates at a faster rate than [RuIII(NH3)5(dIno)] owing to the stronger electron-withdrawing effect of exocyclic oxygen in Xao. The activation parameters ??H ? and ??S ? for k 1 of [RuIII(NH3)5(dIno)] were 80.2?±?15.2?kJ?mol?1 and 47.6?±?9.8?J?K?1 mol?1, respectively, indicating that the disproportionation of RuIII to RuII and RuIV is favored owing to the positive entropy of activation. The final products of both complexes in basic solution under Ar were compared with those under O2. Under both conditions [Ru(NH3)5(8-oxo-L)] was produced, but via different mechanisms. In both aerobic and anaerobic conditions, the deprotonation of highly positively polarized C8-H of Ru-L by OH? initiates a two-electron redox reaction. For the next step, we propose a one-step two-electron redox reaction between L and RuIV under anaerobic conditions, which differentiates from Clarke??s mechanism of two consecutive one-electron redox reactions between L, RuIII, and O2.  相似文献   

17.
Glutathione-dependent enzymes play important protective, repair, or metabolic roles in cells. In particular, enzymes in the glutathione S-transferase (GST) superfamily function in stress responses, defense systems, or xenobiotic detoxification. Here, we identify novel features of bacterial GSTs that cleave β-aryl ether bonds typically found in plant lignin. Our data reveal several original features of the reaction cycle of these GSTs, including stereospecific substrate recognition and stereoselective formation of β-S-thioether linkages. Products of recombinant GSTs (LigE, LigP, and LigF) are β-S-glutathionyl-α-keto-thioethers that are degraded by a β-S-thioetherase (LigG). All three Lig GSTs produced the ketone product (β-S-glutathionyl-α-veratrylethanone) from an achiral side chain-truncated model substrate (β-guaiacyl-α-veratrylethanone). However, when β-etherase assays were conducted with a racemic model substrate, β-guaiacyl-α-veratrylglycerone, LigE- or LigP-catalyzed reactions yielded only one of two potential product (β-S-glutathionyl-α-veratrylglycerone) epimers, whereas the other diastereomer (differing in configuration at the β-position (i.e. its β-epimer)) was produced only in the LigF-catalyzed reaction. Thus, β-etherase catalysis causes stereochemical inversion of the chiral center, converting a β(R)-substrate to a β(S)-product (LigE and LigP), and a β(S)-substrate to a β(R)-product (LigF). Further, LigG catalyzed glutathione-dependent β-S-thioether cleavage with β-S-glutathionyl-α-veratrylethanone and with β(R)-configured β-S-glutathionyl-α-veratrylglycerone but exhibited no or significantly reduced β-S-thioether-cleaving activity with the β(S)-epimer, demonstrating that LigG is a stereospecific β-thioetherase. We therefore propose that multiple Lig enzymes are needed in this β-aryl etherase pathway in order to cleave the racemic β-ether linkages that are present in the backbone of the lignin polymer.  相似文献   

18.
A series of mononuclear manganese(III) complexes of formulae [Mn(L)(X)(H2O)] (1-13) and [Mn(L)(X)] (14-17) (X = ClO4, F, Cl, Br, I, NCS, N3), derived from the Schiff bases of 5-bromosalicylaldehyde and different types of diamine (1,2-diaminoethane, 1,2-diaminopropane, 1,3-diaminopropane and 1,4-diaminobutane), have been synthesized and characterized by the combination of IR, UV-Vis spectroscopies, cyclic voltammetry and by X-ray crystallography. The redox properties of all the manganese(III) complexes show grossly identical features consisting of a reversible or quasireversible MnIII/MnII reduction. Besides MnIII/MnII reduction, the complexes 4, 5, 10, 13 and 16 also show reversible or quasireversible MnIII/MnIV oxidation. A linear correlation has been found for the complexes 5, 7, 11 and 13 [Mn(L2)(X)(H2O)] (X = F, Cl, Br, I) when E1/2 [MnIII/MnII] is plotted against Mulliken electronegativities (χM). The effect of the flexibility of the ligand on redox potential has been studied. It has been observed that the manganese(II) state is stabilized with increasing flexibility of the ligand environment. The crystal structure of 6 shows an octahedral geometry.  相似文献   

19.
Herein, we report the design and synthesis of the novel 12-membered non-antibiotic macrolide (8R,9S)-8,9-dihydro-6,9-epoxy-8,9-anhydropseudoerythromycin A (EM900), which was found to be a potent anti-inflammatory and/or immunomodulatory agent, capable of promoting monocyte to macrophage differentiation. This molecule shows improved acid stability, does not exhibit any anti-bacterial activity and has relatively low cytotoxicity against THP-1 cells. In addition, one of its analogues, (8R,9S)-4″,13-O-diacetyl-8,9-dihydro-6,9-epoxy-8,9-anhydropseudoerythromycin A (EM911), was found to be twice as effective as EM900.  相似文献   

20.
Reactions of 2-(arylazo)aniline, HL [H represents the dissociable protons upon orthometallation and HL is p-RC6H4N = NC6H4-NH2; R = H for HL1; CH3 for HL2 and Cl for HL3] with Ru(R1-tpy)Cl3 (where R1-tpy is 4′-(R1)-2,2′,6′′,2′′-terpyridine and R1 = H or 4-N,N-dimethylaminophenyl or 4-methylphenyl) afford a group of complexes of type [Ru(L)(R1-tpy)]·ClO4 each of which contains C,N,N coordinated L as a tridentate ligand along with a terpyridine. Structure of one such complex has been determined by X-ray crystallography. All the Ru(II) complexes are diamagnetic, display characteristic 1H NMR signals and intense dπ(RuII) → π∗(tpy) MLCT transitions in the visible region. Cyclic voltammetric studies on [Ru(L)(R1-tpy)]·ClO4 complexes show Ru(II)-Ru(III) oxidation within 0.63-0.67 V versus SCE.  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