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 515 毫秒
1.
The optically active mixed‐ligand fac(S)‐tris(thiolato)rhodium(III) complexes, ΔLfac(S)‐[Rh(aet)2(L‐cys‐N,S)]? (aet = 2‐aminoethanethiolate, L‐cys = L‐cysteinate) ( 1 ) and ΔLLfac(S)‐[Rh(aet)(L‐cys‐N,S)2]2? were newly prepared by the equatorial preference of the carboxyl group in the coordinated L‐cys ligand. The amide formation reaction of 1 with 1,10‐diaminodecane and polyallylamine gave the diamine‐bridged dinuclear Rh(III) complex and the single‐chain polymer‐supported Rh(III) complex with retention of the ΔL configuration of 1 , respectively. These Rh(III) complexes reacted with Co(III) or Co(II) to give the linear‐type trinuclear structure with the S‐bridged Co(III) center and the two Δ‐Rh(III) terminal moieties. The polymer‐supported Rh(III) complex was applied not only to the CD spectropolarimetric detection and determination of a trace of precious metal ions such as Au(III), Pt(II), and Pd(II) but also to concentration and extraction of these metal ions into the solid polymer phase. Chirality 28:85–91, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

2.
Summary Recent voltammetric analyses indicate that Shewanella putrefaciens strain 200 produces soluble organic‐Fe(III) complexes during anaerobic respiration of sparingly soluble Fe(III) oxides. Results of the present study expand the range of Shewanella species capable of producing soluble organic‐Fe(III) complexes to include Shewanella oneidensis MR‐1. Soluble organic‐Fe(III) was produced by S. oneidensis cultures incubated anaerobically with Fe(III) oxides, or with Fe(III) oxides and the alternate electron acceptor fumarate, but not in the presence of O2, nitrate or trimethylamine‐N‐oxide. Chemical mutagenesis procedures were combined with a novel MicroElectrode Screening Array (MESA) to identify four (designated Sol) mutants with impaired ability to produce soluble organic‐Fe(III) during anaerobic respiration of Fe(III) oxides. Two of the Sol mutants were deficient in anaerobic growth on both soluble Fe(III)‐citrate and Fe(III) oxide, yet retained the ability to grow on a suite of seven alternate electron acceptors. The rates of soluble organic‐Fe(III) production were proportional to the rates of iron reduction by the S. oneidensis wild‐type and Sol mutant strains, and all four Sol mutants retained wild‐type siderophore production capability. Results of this study indicate that the production of soluble organic‐Fe(III) may be an important intermediate step in the anaerobic respiration of both soluble and sparingly soluble forms of Fe(III) by S. oneidensis.  相似文献   

3.
An N‐butylamine functionalized graphene oxide nanolayer was synthesized and characterized by ultraviolet (UV)–visible spectrometry, Fourier transform infrared spectroscopy, X‐ray photoelectron spectroscopy, and transmission electron microscopy. Detection of iron(III) based on photoluminescence spectroscopy was investigated. The N‐butylamine functionalized graphene oxide was shown to specifically interact with iron (III), compared with other cationic trace elements including potassium (I), sodium (I), calcium (II), chromium (III), zinc (II), cobalt (II), copper (II), magnesium (II), manganese (II), and molybdenum (VI). The quenching effect of iron (III) on the luminescence emission of N‐butylamine functionalized graphene oxide layer was used to detect iron (III). The limit of detection (2.8 × 10?6 M) and limit of quantitation (2.9 × 10?5 M) were obtained under optimal conditions. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
We developed N,N′-bis(salicylidene)-1,2-phenylenediamine (salophene, 1) as a chelating agent for metal ions such as Mn(II/III), Fe(II/III), Co(II), Ni(II), Cu(II), and Zn(II). The resulting complexes, from which owing to the carrier ligand a selective mode of action is assumed, were tested for antiproliferative effects on the MCF-7 breast cancer cell line. The cytotoxicity in this assay depended on the nature of the transition metal used. Iron complexes in oxidation states +II and +III (3, 4) strongly reduced cell proliferation in a concentration-dependent manner, whereas, e.g., the manganese analogues 5 and 6 were only marginally active. Therefore, the [N,N′-bis(salicylidene)-1,2-phenylenediamine]iron(II/III) complexes 3 and 4 were selected for studies on the mode of action. Both complexes possessed high activity against various tumor cells, for instance, MDA-MB-231 mammary carcinoma cells as well as HT-29 colon carcinoma cells. They were able to generate reactive oxygen species, showed DNA binding, and induced apoptosis. Exchange of 1 by N,N′-bis(salicylidene)-1,2-cyclohexanediamine (saldach, 2) yielding complexes 7 and 8 reduced the in vitro effects drastically. An unequivocal mode of action cannot be deduced from these results, but it seems to be very likely that cell death is caused by interference with more than one intracellular target. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

5.
2‐(1H‐Tetrazol‐5‐yl)pyridine ( L ) has been reacted separately with Me2NCH2CH2Cl?HCl and ClCH2CH2OH to yield two regioisomers in each case, N,N‐dimethyl‐2‐[5‐(pyridin‐2‐yl)‐1H‐tetrazol‐1‐yl]ethanamine ( L1 )/N,N‐dimethyl‐2‐[5‐(pyridin‐2‐yl)‐2H‐tetrazol‐2‐yl]ethanamine ( L2 ) and 2‐[5‐(pyridin‐2‐yl)‐1H‐tetrazol‐1‐yl]ethanol ( L3 )/2‐[5‐(pyridin‐2‐yl)‐2H‐tetrazol‐2‐yl]ethanol ( L4 ), respectively. These ligands, L1 – L4 , have been coordinated with CuCl2?H2O in 1 : 1 composition to furnish the corresponding complexes 1 – 4 . EPR Spectra of Cu complexes 1 and 3 were characteristic of square planar geometry, with nuclear hyperfine spin 3/2. Single X‐ray crystallographic studies of 3 revealed that the Cu center has a square planar structure. DNA binding studies were carried out by UV/VIS absorption; viscosity and thermal denaturation studies revealed that each of these complexes are avid binders of calf thymus DNA. Investigation of nucleolytic cleavage activities of the complexes was carried out on double‐stranded pBR322 circular plasmid DNA by using a gel electrophoresis experiment under various conditions, where cleavage of DNA takes place by oxidative free‐radical mechanism (OH ? ). In vitro anticancer activities of the complexes against MCF‐7 (human breast adenocarcinoma) cells revealed that the complexes inhibit the growth of cancer cells. The IC50 values of the complexes showed that Cu complexes exhibit comparable cytotoxic activities compared to the standard drug cisplatin.  相似文献   

6.
The synthetic, spectroscopic, and biological studies of Cu(II), Ni(II), Zn(II), Co(II), Mn(II), Fe(III) and Cr(III) complexes of N4-(7′-chloroquinoline-4′-ylamino)-N1-(2-hydroxy-benzylidene)thiosemicarbazone (HL) obtained by the reaction of N4-(7′-chloroquinolin-4′-ylamino)thiosemicarbazide with 2-hydroxybenzaldehyde. The structures of the complexes were determined on the basis of the elemental analyses, spectroscopic data (IR, electronic, 1H and 13C NMR and Mass spectra) along with magnetic susceptibility measurements, molar conductivity and thermogravimetric analyses. Electrical conductance measurement revealed the non-electrolytic nature of the complexes. The resulting colored products are mononuclear in nature. On the basis of the above studies, only one ligand was suggested to be coordinated to each metal atom by thione sulfur, azomethine nitrogen and phenolic oxygen to form mononuclear complexes in which the thiosemicarbazone behaves as a monobasic tridendate ligand. The ligand and its metal complexes were tested against Gram + ve bacteria (Staphylococcus aureus), Gram ? ve bacteria (Escherichia coli), fungi (Candida albicans) and (Fusarium solani). The tested compounds exhibited significant activity.  相似文献   

7.
A novel series of complexes of the type [M(C36H22N6)X]X2, where M = Cr(III), Mn(III), Fe(III); X = Cl?, NO3?, CH3COO?; and (C36H22N6) corresponds to the tetradentate macrocyclic ligand, have been synthesized by condensation of 1,8-diaminonaphthalene and isatin in the presence of trivalent metal salts in methanolic medium. The complexes have been characterized by elemental analysis, conductance and magnetic measurements, and UV/Vis, IR, and mass spectroscopy. On the basis of these studies, a five coordinate square pyramidal geometry for all of these complexes is proposed. All synthesized macrocyclic complexes have been tested for in vitro antimicrobial activities against some pathogenic bacterial strains, viz. Staphylococcus aureus, Bacillus subtilis (Gram-positive), Escherichia coli, Pseudomonas aeruginosa (Gram-negative), and two fungal strains, viz. Aspergillus niger, Aspergillus flavus. The MICs shown by the complexes against these microbial strains have been compared with MICs shown by standard antibiotic ciprofloxacin and the antifungal drug amphotericin-B.  相似文献   

8.
We have previously reported on a gold(III) complex, namely [AuBr2(DMDT)] (N,N‐dimethyldithiocarbamate) showing potent in vitro and in vivo growth inhibitory activities toward human cancer cells and identifying the cellular proteasome as one of the major targets. However, the importance of the oxidation state of the gold center and the involved mechanism of action has yet to be established. Here we show that both gold(III)? and gold(I)–dithiocarbamato species, namely [AuBr2(ESDT)] (AUL12) and [Au(ESDT)]2 (AUL15), could inhibit the chymotrypsin‐like activity of purified 20S proteasome and 26S proteasome in human breast cancer MDA‐MB‐231 cells, resulting in accumulation of ubiquitinated proteins and proteasome target proteins, and induction of cell death, but at significantly different levels. Gold(I)‐ and gold(III)‐compound‐mediated proteasome inhibition and cell death induction were completely reversed by the addition of a reducing agent, dithiothreitol or N‐acetyl‐L ‐cysteine, suggesting the involvement of redox processes. Furthermore, treatment of MDA‐MB‐231 cells with gold(III) compound (AUL12), but not the gold(I) analog (AUL15), resulted in the production of significant levels of reactive oxygen species. Our study provides strong evidence that the cellular proteasome is an important target of both gold(I) and gold(III)–dithiocarbamates, but distinct cellular mechanisms of action are responsible for their different overall effect. J. Cell. Biochem. 109: 162–172, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

9.
The spectroscopy, electrochemistry and electrogenerated chemiluminescence (ECL) of eight bisalicylideneethylenediamino (salen) metal complexes are reported. Two of the complexes contain an unsubstituted salen ligand and either cobalt(II) or nickel(II). The others have 1,2-cyclohexanediamonio-N,N′-bis(3,5-di-t-butylsalicylidene) as the ligand, and chromium(III), aluminum(III), cobalt(II), cobalt(III) or manganese(II) as the metal center. The complexes have lowest energy absorption maxima between 350 and 430 nm. When excited at these wavelengths, the complexes emit between 417 and 594 nm in acetonitrile. Photoluminescence efficiencies (?em) were between 0.0310 and 23.8 compared to Ru(bpy)32+ (bpy = 2,2′-bipyridine; ?em = 1), with the aluminum complexes displaying the most intense photoluminescence. Both reversible and irreversible oxidative electrochemistry is displayed by the metal–salen complexes with oxidation potentials ranging between +0.152 and +1.661 V versus Ag/AgCl. The ECL intensity peaks at a potential corresponding to oxidation of both TPrA and the salen systems, indicating that both are involved in the ECL reaction sequence. ECL efficiencies (?ecl) were between 0.0018 and 0.0086 when compared to Ru(bpy)32+ (?ecl = 1) in acetonitrile (0.05 M tri-n-propylamine (TPrA) as an oxidative–reductive ECL coreactant). Also, qualitative studies using transmission filters suggest that the complexes emit ECL in approximately the same region as their photoluminescence, indicating that the same excited state is formed in both experiments.  相似文献   

10.
To explore the relationship between the structure of the ligands and the luminescent properties of the lanthanide complexes, luminescent lanthanide complexes of a new tripodal ligand, featuring N‐thenylsalicylamide arms, were synthesized and characterized by elemental analysis, IR and TGA measurements. Photophysical properties of the complexes were studied by means of UV ? visible absorption and steady‐state luminescence spectroscopy. The results of UV ? vis spectra indicate that metal binding does not disturb the electronic structure of the ligand. Excited‐state luminescence lifetimes and quantum yields of the complexes were determined. The photoluminescence analysis suggested that there is an efficient ligand ? Ln(III) energy transfer for the Tb(III) complex, and the ligand is an efficient 'antenna' for Tb(III). From a more general perspective, the results demonstrated the potential application of the lanthanide complex as luminescent materials in material chemistry. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
《Chirality》2017,29(6):273-281
Enantiomeric 1H and 13C NMR signal separation behaviors of various α‐amino acids and DL‐tartarate were investigated by using the samarium(III) and neodymium(III) complexes with (S ,S )‐ethylenediamine‐N ,N' ‐disuccinate as chiral shift reagents. A relatively smaller concentration ratio of the lanthanide(III) complex to substrates was suitable for the neodymium(III) complex compared with the samarium(III) one, striking a balance between relatively greater signal separation and broadening. To clarify the difference in the signal separation behavior, the chemical shifts of β‐protons for fully bound D‐ and L‐alanine (δb(D) and δb(L)) and their adduct formation constants (K s) were obtained for both metal complexes. Preference for D‐alanine was similarly observed for both complexes, while it was revealed that the difference between the δb(D) and δb(L) values is the significant factor to determine the enantiomeric signal separation. The neodymium(III) and samarium(III) complexes can be used complementarily for higher and smaller concentration ranges of substrates, respectively, because the neodymium(III) complex gives the larger difference between the δb(D) and δb(L) values with greater signal broadening compared to the samarium(III) complex.  相似文献   

12.
Salicylidenimine palladium(II) complexes trans‐Pd(O,N)2 adopt step and bowl arrangements. A stereochemical analysis subdivides 52 compounds into 41 step and 11 bowl types. Step complexes with chiral N‐substituents and all the bowl complexes induce chiral distortions in the square planar system, resulting in Δ/Λ configuration of the Pd(O,N)2 unit. In complexes 1 , 2 , 3 , 4 , 5 , 6 with enantiomerically pure N‐substituents ligand chirality entails a specific square chirality and only one diastereomer assembles in the lattice. Dimeric Pd(O,N)2 complexes with bridging N‐substituents in trans‐arrangement are inherently chiral. For dimers 7 , 8 , 9 , 10 , 11 different chirality patterns for the Pd(O,N)2 square are observed. The crystals contain racemates of enantiomers. In complex 12 two independent molecules form a tight pair. The (RC) configuration of the ligand induces the same Δ chirality in the Pd(O,N)2 units of both molecules with varying square chirality due to the different crystallographic location of the independent molecules. In complexes 13 and 14 atrop isomerism induces specific configurations in the Pd(O,N)2 bowl systems. The square chirality is largest for complex 15 [(Diop)Rh(PPh3)Cl)], a catalyst for enantioselective hydrogenation. In the lattice of 15 two diastereomers with the same (RC,RC) configuration in the ligand Diop but opposite Δ and Λ square configurations co‐crystallize, a rare phenomenon in stereochemistry. Chirality 25:663–667, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Pyochelin, its analog 3′′-nor-NH-pyochelin, and the related methyl hydroxamate, 2-(2′-hydroxyphenyl)-4,5-dihydrothiazol-4-carboxylic acid methoxymethyl amide, have been prepared together with their Fe(III) complexes. The solution stoichiometry and the coordination of the three Fe(III) complexes in methanol or buffered (pH∼2) 50:50 (v/v) methanol–water mixtures were determined using various spectroscopic methods: UV–vis absorption, X-ray absorption, extended X-ray absorption fine structure and electron paramagnetic resonance. All three systems showed both a 1:1 and 2:1 ligand–Fe(III) stoichiometry, but presented different coordination properties. Conditional formation constants (pH∼2) were determined for both the 1:1 and 2:1 complexes in all three systems. Computation of the coordination-conformational energies by semiempirical methods indicated that the coordination in the case of the 2:1 complexes of pyochelin–Fe(III) and 3′′-nor-NH-pyochelin–Fe(III) was asymmetrical, with one molecule of pyochelin (or 3′′-nor-NH-pyochelin) tetradentately coordinated (O1, N1, N2 and O3) to the Fe(III), and the second molecule bound bidentately (O1, N1 or N2, O3), to complete the octahedral geometry. In contrast, two molecules of the methyl hydroxamate each provided a set of tridentate ligand atoms in the formation of the 2:1 ligand–Fe(III) complex. These results are consistent with the role of pyochelin in the uptake of iron by the FptA receptor in the outer membrane of Pseudomonas aeruginosa and in several gram-negative bacteria.  相似文献   

14.
A new class of polydentate Mannich bases featuring an N2S2 donor system, bis((2-mercapto-N-phenylacetamido)methyl)phosphinic acid H3L1 and bis((2-mercapto-N-propylacetamido)methyl)phosphinic acid H3L2, has been synthesised from condensation of phosphinic acid and paraformaldehyde with 2-mercaptophenylacetamide W1 and 2-mercaptopropylacetamide W2, respectively. Monomeric complexes of these ligands, of general formula K2[CrIII(Ln)Cl2], K3[M′II(Ln)Cl2] and K[M(Ln)] (M′ = Mn(II) or Fe(II); M = Co(II), Ni(II), Cu(II), Zn(II), Cd(II) or Hg(II); n = 1, 2) are reported. The structures of new ligands, 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) and Fe(II) complexes, square planar for Ni(II) and Cu(II) complexes and tetrahedral for the Co(II), Zn(II), Cd(II) and Hg(II) complexes. Complex formation studies via molar ratio in DMF solution were consistent to those found in the solid complexes with a ratio of (M:L) as (1:1).  相似文献   

15.
Reaction of the potent hydroxamate-based histone deacetylase (HDAC) inhibitor, suberoylanilide hydroxamic acid (SAHA), with hydrated metal salts of Fe(III), Cu(II), Ni(II) and Zn(II) yielded a tris-hydroxamato complex in the case of Fe(III) and bis-hydroxamato complexes in the case of Cu(II), Ni(II) and Zn(II) both in the solid state and in solution. Reaction of the secondary hydroxamic acid, N-Me-SAHA, also yielded a tris-hydroxamato complex in the case of Fe(III) and bis-hydroxamato complexes in the case of Cu(II), Ni(II) and Zn(II) in solution. These metal complexes have the hydroxamato moiety coordinated in an O,O’-bidentate fashion. Stability constants of the metal complexes formed with SAHA and N-Me-SAHA in a DMSO/H2O 70/30%(v/v) mixture are described. A novel crystal structure of SAHA together with a novel synthesis for N-Me-SAHA are also reported.  相似文献   

16.
Iron acquisition by iron‐limited cyanobacteria is typically considered to be mediated mainly by siderophores, iron‐chelating molecules released by iron‐limited cyanobacteria into the environment. In this set of experiments, iron uptake by iron‐limited cells of the cyanobacterium Anabaena flos‐aquae (L.) Bory was investigated in cells resuspended in siderophore‐free medium. Removal of siderophores decreased iron‐uptake rates by ~60% compared to siderophore‐replete conditions; however, substantial rates of iron uptake remained. In the absence of siderophores, Fe(III) uptake was much more rapid from a weaker synthetic chelator [N‐(2‐hydroxyethyl)ethylenediamine‐N,N′,N′‐triacetic acid (HEDTA); log Kcond = 28.64 for Fe(III)HEDTA(OH)?] than from a very strong chelator [N,N′‐bis(2‐hydroxybenzyl)‐ethylenediamine‐N,N′‐diacetic acid (HBED); log Kcond = 31.40 for Fe(III)HBED?], and increasing chelator:Fe(III) ratios decreased the Fe(III)‐uptake rate; these results were evident in both short‐term (4 h; absence of siderophores) and long‐term (116 h; presence of siderophores) experiments. However, free (nonchelated) Fe(III) provided the most rapid iron uptake in siderophore‐free conditions. The results of the short‐term experiments are consistent with an Fe(III)‐binding/uptake mechanism associated with the cyanobacterial outer membrane that operates independently of extracellular siderophores. Iron uptake was inhibited by temperature‐shock treatments of the cells and by metabolically compromising the cells with diphenyleneiodonium; this finding indicates that the process is dependent on active metabolism to operate and is not simply a passive Fe(III)‐binding mechanism. Overall, these results point to an important, siderophore‐independent iron‐acquisition mechanism by iron‐limited cyanobacterial cells.  相似文献   

17.
A series of N,N′‐dibisphosphonate‐containing 1,3‐propanediamine derivatives ( L1 – L6 ) and their corresponding dichloridoplatinum(II) complexes ( 1 – 6 ) have been synthesized and characterized by elemental analysis, 1H‐NMR, 13C‐NMR, 31P‐NMR and HR‐MS spectra. The in vitro antitumor activities of compounds L1 – L6 and 1 – 6 were tested by WST‐8 assay with Cell Counting Kit‐8, indicating that platinum‐based complexes 1 – 6 showed higher cytotoxicity than corresponding ligands L1 – L6 against A549 and MG‐63, especially complex 2 which displayed comparable cytotoxicity to those of cisplatin and zoledronate after 48 h incubation. In addition, complexes 1 – 6 were more active in vitro on osteosarcoma cell line MG‐63 than normal osteoblast cell line hFOB 1.19. The structure‐activity relationship has been summarized based on the in vitro cytotoxicity of three series of platinum complexes from this and our previous studies. The in vitro bone affinity of platinum complexes was also tested by hydroxyapatite (HAP) chromatography in terms of capacity factor K′. Besides, in this paper, representative complex 2 , which has been proved to be a promising antitumor agent with high cytotoxicity and bone HAP binding property, was investigated for its mechanism of action producing cell death against MG‐63.  相似文献   

18.
The synthesis of poly[N,N‐bis(3‐aminopropyl)glycine] (PAPGly) dendrons Gd‐based contrast agents (GdCAs) via an orthogonal protection of the different functional groups and an activation/coupling strategy wherein a specific number of synthetic steps add a generation to the existing dendron has been described. The aim of this protocol is to build up two different generations of dendrons ( G‐0 or dendron's core, and G‐1 ) with peripheral NH2 groups to conjugate a 1,4,7,10‐tetraazacyclododecane‐1,4,7‐triacetic acid (DO3A) derivative and afterwards to chelate with Gd3+ paramagnetic ions. These complexes, which have a well‐defined molecular weight, are of relevance to MRI as an attempt to gain higher 1H relaxivity by slowing down the rotation of molecule compared to monomeric Gd(III) complexes used as contrast agents and to increase the number of paramagnetic centers present in one molecular structure. From the study of their water 1H longitudinal relaxation rate at different magnetic fields (NMRD, Nuclear Magnetic Relaxation Dispersion) and by evaluating the variable temperature 17O‐NMR data we determined the parameters characterizing the water exchange rate and the rotational correlation time of each complex, both affecting 1H relaxivity. Furthermore, these two novel PAPGly GdCAs were objects of i) an in vivo study to determine their biodistributions in healthy C57 mice at several time points, and ii) the Dynamic Contrast‐Enhanced MRI (DCE‐MRI) approach to assess their contrast efficiency measured in the tumor region of C57BL/6 mice transplanted subcutaneously with B16‐F10 melanoma cells. The aim of the comparison of these two dendrons GdCAs, having different molecular weights (MW), is to understand how MW and relaxivity may influence the contrast enhancement capabilities in vivo at low magnetic field (1 T). Significant contrast enhancement was observed in several organs (vessel, spleen and liver), already at 5 min post‐injection, for the investigated CAs. Moreover, these CAs induced a marked contrast enhancement in the tumor region, thanks to the enhanced permeability retention effect of those macromolecular structures.  相似文献   

19.
The preparation of tetrakis((+)‐hfbc) lanthanide(III) complexes with an encapsulated alkali metal and ammonium ions M[Ln((+)‐hfbc)4] (hereafter abbreviated as M‐Ln : (+)‐hfbc, (+)‐heptafluorobutyrylcamphorate; M, ammonium or benzyl ammonium ions as well as alkali metal ions) was reported and discussed. The electronic circular dichroism (CD) spectra in the intraligand π?π* transition of M–Ln were examined in view of the solvent effect. Here, the concentration, alkali metal, and ammonium ion dependences are compared with the solid CD, 5D07F0(Eu(III)) excitation spectra, circularly polarized luminescence, and vibrational circular dichroism. It has been revealed that the dodecahedral eight coordinate DD‐8‐M‐Ln complexes in crystals are equilibrated between the diastereoselectively formed square antiprism eight coordinate SAPR‐8‐M‐Ln and [Ln((+)‐hfbc)3] in EtOH and CH3CN solutions or between the SAPR‐8‐M‐Ln and DD‐D2d(mmmm)‐8‐M‐Ln complexes in CHCl3 solution. The observed CD couplets are found to reflect the exciton CD couplets which are useful to determine the four‐bladed SAPR‐(llll) absolute configuration around the lanthanide(III) ion. Chirality 24:1055–1062, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

20.
The sequential action of five distinct endosomal‐sorting complex required for transport (ESCRT) complexes is required for the lysosomal downregulation of cell surface receptors through the multivesicular body (MVB) pathway. On endosomes, the assembly of ESCRT‐III is a highly ordered process. We show that the length of ESCRT‐III (Snf7) oligomers controls the size of MVB vesicles and addresses how ESCRT‐II regulates ESCRT‐III assembly. The first step of ESCRT‐III assembly is mediated by Vps20, which nucleates Snf7/Vps32 oligomerization, and serves as the link to ESCRT‐II. The ESCRT‐II subunit Vps25 induces an essential conformational switch that converts inactive monomeric Vps20 into the active nucleator for Snf7 oligomerization. Each ESCRT‐II complex contains two Vps25 molecules (arms) that generate a characteristic Y‐shaped structure. Mutant ‘one‐armed’ ESCRT‐II complexes with a single Vps25 arm are sufficient to nucleate Snf7 oligomerization. However, these oligomers cannot execute ESCRT‐III function. Both Vps25 arms provide essential geometry for the assembly of a functional ESCRT‐III complex. We propose that ESCRT‐II serves as a scaffold that nucleates the assembly of two Snf7 oligomers, which together are required for cargo sequestration and vesicle formation during MVB sorting.  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