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 140 毫秒
1.
2.
As part of an overall effort to map the energetic landscape of the base excision repair pathway, we report the first thermodynamic characterization of repair enzyme binding to lesion-containing duplexes. Isothermal titration calorimetry (ITC) in conjunction with spectroscopic measurements and protease protection assays have been employed to characterize the binding of Escherichia coli formamidopyrimidine-glycosylase (Fpg), a bifunctional repair enzyme, to a series of 13-mer DNA duplexes. To resolve energetically the binding and the catalytic events, several of these duplexes are constructed with non-hydrolyzable lesion analogs that mimic the natural 8-oxo-dG substrate and the abasic-like intermediates. Specifically, one of the duplexes contains a central, non-hydrolyzable, tetrahydrofuran (THF) abasic site analog, while another duplex contains a central, carbocyclic substrate analog (carba-8-oxo-dG). ITC-binding studies conducted between 5.0 °C and 15.0 °C reveal that Fpg association with the THF-containing duplex is characterized by binding free energies that are relatively invariant to temperature (ΔG∼−9.5 kcal mol−1), in contrast to both the reaction enthalpy and entropy that are strongly temperature-dependent. Complex formation between Fpg and the THF-containing duplex at 15 °C exhibits an unfavorable association enthalpy that is compensated by a favorable association entropy (TΔS=+17.0 kcal mol−1). The entropic nature of the binding interaction, coupled with the large negative heat capacity is consistent with Fpg complexation to the THF-containing duplex involving significant burial of non-polar surface areas. By contrast, under the high ionic strength buffer conditions employed herein (200 mM NaCl), no appreciable Fpg affinity for the carba-8-oxo-dG substrate analog is detected. Our results suggest that initial Fpg recognition of a damaged DNA site is predominantly electrostatic in nature, and does not involve large contact interfaces. Subsequent base excision presumably facilitates accommodation of the resulting lesion site into the binding pocket, as the enzyme interaction with the THF-containing duplex is characterized by high affinity and a large negative heat capacity change. Our data are consistent with a pathway in which Fpg glycosylase activity renders the base excision product a preferred ligand relative to the natural substrate, thereby ensuring the fidelity of removing highly reactive and potentially mutagenic abasic-like intermediates through catalytic elimination reactions.  相似文献   

3.
Nucleotide excision repair (NER) is a vital cellular defense system against carcinogen-DNA adducts, which, if not repaired, can initiate cancer development. The structural features of bulky DNA lesions that account for differences in NER efficiencies in mammalian cells are not well understood. In vivo, the predominant DNA adduct derived from metabolically activated benzo[a]pyrene (BP), a prominent environmental carcinogen, is the 10S (+)-trans-anti-[BP]-N2-dG adduct (G*), which resides in the B-DNA minor groove 5′-oriented along the modified strand. We have compared the structural distortions in double-stranded DNA, imposed by this adduct, in the different sequence contexts 5′-…CGG*C…, 5′-…CG*GC…, 5′-…CIG*C… (I is 2′-deoxyinosine), and 5′-…CG*C…. On the basis of electrophoretic mobilities, all duplexes manifest moderate bends, except the 5′-…CGG*C…duplex, which exhibits an anomalous, slow mobility attributed to a pronounced flexible kink at the site of the lesion. This kink, resulting from steric hindrance between the 5′-flanking guanine amino group and the BP aromatic rings, both positioned in the minor groove, is abolished in the 5′-…CIG*C…duplex (the 2′-deoxyinosine group, I, lacks this amino group). In contrast, the sequence-isomeric 5′-…CG*GC…duplex exhibits only a moderate bend, but displays a remarkably increased opening rate at the 5′-flanking base pair of G*, indicating a significant destabilization of Watson-Crick hydrogen bonding. The NER dual incision product yields were compared for these different sequences embedded in otherwise identical 135-mer duplexes in cell-free human HeLa extracts. The yields of excision products varied by a factor of as much as ∼ 4 in the order 5′-...CG*GC…> 5′...CGG*C…≥ 5′...CIG*C…≥ 5′-…CG*C…. Overall, destabilized Watson-Crick hydrogen bonding, manifested in the 5′-...CG*GC...duplex, elicits the most significant NER response, while the flexible kink displayed in the sequence-isomeric 5′-...CGG*C...duplex represents a less significant signal in this series of substrates. These results demonstrate that the identical lesion can be repaired with markedly variable efficiency in different local sequence contexts that differentially alter the structural features of the DNA duplex around the lesion site.  相似文献   

4.
Five new octahedral iron(II) complexes [FeL2(4-dpa)]n(EtOH) (1), [FeL2(bipy)]n(DMF) (2), [FeL1(bpee)]n (3), [Fe2L3(1-meim)4](1-meim)4 (4) and [FeL1(DMAP)2] (5), with L1 and L2 being tetradentate coordinating Schiff base like ligands (L1 = (E,E)-[{diethyl-2,2′-[1,2-phenylenebis(iminomethylidyne)]bis[3-oxobutanato](2-)-N,N′,O3,O3′}, L2 = (3,3′)-[{1,2-phenylenebis(iminomethylidyne)]bis(2,4-pentane-dionato)(2-)-N,N′,O2,O2′}) and L3 being a octadentate dinucleating coordinating Schiff base like ligand ({tetraethyl-(E,E,E,E)-2,2′,2′′,2′′′-[1,2,4,5-phenylentetra(iminomethylidine)]tetra[3-oxobutanoato](2-)-N,N′,N′′,N′′′,O3,O3′,O3′′,O3′′′}); 4-dpa = di(4-picolyl)-amine, bipy = 4,4′-bipyridine, bpee = trans-1,2-bis(4-pyridyl)ethylene, 1-meim = 1-methylimidazole and DMAP = 4-dimethylaminopyridine, have been synthesized and characterised using X-ray structure analysis and T-dependent susceptibility measurements. Both methods indicate that all iron(II) centres are in the paramagnetic high-spin state over the whole temperature range investigated. The O-Fe-O angle, the so called bit of the equatorial ligand, is with an average of 111° in the region typical for high-spin iron(II) complexes of this ligand type. In the case of compound 1 an infinite two-dimensional hydrogen bond network can be found, for the compounds 2-4 no hydrogen bond interactions are observed between the complex molecules. A comparison of the curve progression obtained from the magnetic measurements of the mononuclear complex 5 and the polymeric complexes 1-3 leads to the conclusion that no magnetic interactions are mediated over the bridging axial ligands. For the dinuclear complex 4 weak antiferromagnetic interactions between the two iron centres are found.  相似文献   

5.
Spectroscopic (IR, 31P NMR and UV-Vis) and electrochemical studies on fac-[Mn(CO)3(L′-L′)(L)]0/+,where L′-L′ = 1,2-bis(diphenylphosphino)ethane (dppe) or 1,10-phenanthroline (phen) and L = bromide, triflate, imidazole (im), isonicotinamide (isn) or N-(2-hydroxyethyl)isonicotinamide (heisn), were undertaken to understand the effect of various ligands on the CO-Mn-L and CO-Mn-(L′-L′) bonding characteristics of these complexes. Crystal structures for L = triflate/L′-L′ = dppe, L = triflate/L′-L′ = phen and L = isn/L′-L′ = phen are reported and they show that the two Mn-O(OSO2CF3) and Mn-N(isn) distances are similar. The tricarbonyl complexes exhibit two major bands in the 250-300 and 350-450 nm region of the UV-Vis spectrum. The lowest energy bands have been assigned as a contribution from both the metal-centered (MC) and metal to ligand (dπ → L′-L′) charge transfer (MLCT) transitions. The energy of this maximum absorption decreases in the order Br ∼ triflate > im > isn ∼ heisn. The cyclic four-component mechanism was observed at room temperature by voltammetric techniques for all the cases. On the basis of d metal orbital splitting, an electronic molecular orbital diagram is proposed. In this model, the ligands along the z-axis play a relevant role in the reverse of the HOMO energies of the fac/mer isomers by stabilizing the metal dz2 orbital relative to dxy in mer-Mn(II).  相似文献   

6.
Among individual cells of the same source and type, the complex shear modulus GG exhibits a large log-normal distribution that is the result of spatial, temporal, and intrinsic variations. Such large distributions complicate the statistical evaluation of pharmacological treatments and the comparison of different cell states. However, little is known about the characteristic features of cell-to-cell variation. In this study, we investigated how this variation depends on the spatial location within the cell and on the actin filament cytoskeleton, the organization of which strongly influences cell mechanics. By mechanically probing fibroblasts arranged on a microarray, via atomic force microscopy, we observed that the standard deviation σ   of GG was significantly reduced among cells in which actin filaments were depolymerized. The parameter σ also exhibited a subcellular spatial dependence. Based on our findings regarding the frequency dependence of σ   of the storage modulus GG, we proposed two types of cell-to-cell variation in GG that arise from the purely elastic and the frequency-dependent components in terms of the soft glassy rheology model of cell deformability. We concluded that the latter inherent cell-to-cell variation can be reduced greatly by disrupting actin networks, by probing at locations within the cell nucleus boundaries distant from the cell center, and by measuring at high loading frequencies.  相似文献   

7.
Gradual solvation of [(4,4′-bpy)ReI(CO)3(dppz)]+ (dppz = dipyridil[3,2-a:2′3′-c]phenazine) by water molecules causes a quenching of the emission in accordance with Perrin’s model of spheres. The calculated radius of the sphere, r = 2.6 ± 0.2 ?, is therefore very close to the distance from the Re center to the oxygen atom of the CO ligands, i.e., l = 2.73 ?. In addition, excited state reactions with TEA produce [(4,4′-bpy)ReI(CO)3(dppz)] and [(4,4′-bpy)ReI(CO)3(dppz)]. This experimental observation is inconsistent with the formation of the products in the lowest lying and emissive dppz-centered 3ππ* excited state. Jablonski schemes based on the participation of excited states other than the lowest 3ππ* excited state are proposed.  相似文献   

8.
Electrochemical oxidation of [RuII(terpy)(sq)(NH3)]+ in neutral water (pH 8.0) at +0.8 V (versus SCE) generated [RuII(terpy)(q)(NH2)]2+ and/or [RuIII(terpy)(sq)(NH2)]2+ (terpy = 2,2′:6′,2′′-terpyridine, sq = 3,5-di-tert-butyl-1,2-semiquinonate, q = 3,5-di-tert-butyl-1,2-benzoquinone), which played roles in hydrogen abstraction and one-electron acceptor in the catalytic oxidation of methanol, ethanol, and 2-propanol affording formaldehyde, acetoaldehyde, and acetone, respectively, under the electrolysis conditions.  相似文献   

9.
The most extensively studied ficins have been isolated from the latex of Ficus glabrata and Ficus carica. However the proteases (ficins) from other species are less known. The purification and characterization of a protease from the latex of Ficus racemosa is reported. The enzyme purified to homogeneity is a single polypeptide chain of molecular weight of 44,500 ± 500 Da as determined by MALDI-TOF. The enzyme exhibited a broad spectrum of pH optima between pH 4.5-6.5 and showed maximum activity at 60 ± 0.5 °C. The enzyme activity was completely inhibited by pepstatin-A indicating that the purified enzyme is an aspartic protease. Far-UV circular dichroic spectra revealed that the purified enzyme contains predominantly β-structures. The purified protease is thermostable. The apparent Tm, (mid point of thermal inactivation) was found to be 70 ± 0.5 °C. Thermal inactivation was found to follow first order kinetics at pH 5.5. Activation energy (Ea) was found to be 44.0 ± 0.3 kcal mol−1. The activation enthalpy (ΔH), free energy change (ΔG) and entropy (ΔS) were estimated to be 43 ± 4 kcal mol−1, −26 ± 3 kcal mol−1 and 204 ± 10 cal mol−1 K−1, respectively. Its enzymatic specificity studied using oxidized B chain of insulin indicates that the protease preferably hydrolyzed peptide bonds C-terminal to glutamate, leucine and phenylalanine (at P1 position). The broad specificity, pH optima and elevated thermal stability indicate the protease is distinct from other known ficins and would find applications in many sectors for its unique properties.  相似文献   

10.
Escherichia coli RecBCD is a highly processive DNA helicase involved in double-strand break repair and recombination that possesses two helicase/translocase subunits with opposite translocation directionality (RecB (3′ to 5′) and RecD (5′ to 3′)). RecBCD has been shown to melt out ∼ 5-6 bp upon binding to a blunt-ended duplex DNA in a Mg2+-dependent, but ATP-independent reaction. Here, we examine the binding of E. coli RecBC helicase (minus RecD), also a processive helicase, to duplex DNA ends in the presence and in the absence of Mg2+ in order to determine if RecBC can also melt a duplex DNA end in the absence of ATP. Equilibrium binding of RecBC to DNA substrates with ends possessing pre-formed 3′ and/or 5′ single-stranded (ss)-(dT)n flanking regions (tails) (n ranging from zero to 20 nt) was examined by competition with a fluorescently labeled reference DNA and by isothermal titration calorimetry. The presence of Mg2+ enhances the affinity of RecBC for DNA ends possessing 3′ or 5′-(dT)n ssDNA tails with n < 6 nt, with the relative enhancement decreasing as n increases from zero to six nt. No effect of Mg2+ was observed for either the binding constant or the enthalpy of binding (ΔHobs) for RecBC binding to DNA with ssDNA tail lengths, n ≥ 6 nucleotides. Upon RecBC binding to a blunt duplex DNA end in the presence of Mg2+, at least 4 bp at the duplex end become accessible to KMnO4 attack, consistent with melting of the duplex end. Since Mg2+ has no effect on the affinity or binding enthalpy of RecBC for a DNA end that is fully pre-melted, this suggests that the role of Mg2+ is to overcome a kinetic barrier to melting of the DNA by RecBC and presumably also by RecBCD. These data also provide an accurate estimate (ΔHobs = 8 ± 1 kcal/mol) for the average enthalpy change associated with the melting of a DNA base-pair by RecBC.  相似文献   

11.
Synthesis of three acyclic chiral phosphites is reported, in the form of dithymidine phosphite triesters. These diastereomerically pure P-stereogenic phosphites undergo epimerization at a measurable rate at 150 °C. When the alcohols on the deoxyribose moieties are protected as acyls, decomposition is minimized and by computer fitting, rate constants for epimerization can be extracted. These allow for the first time calculation of the barrier to inversion of configuration in phosphite triesters, giving ΔG(150 °C) = 33.0 ± 0.2 kcal mol−1, comparable to the inversion barrier seen for phosphines.  相似文献   

12.
The new diiron alkynyl methoxy carbene complexes [Fe2{μ-CN(Me)(R)}(μ-CO)(CO){C(OMe)CCR′}(Cp)2]+ (R = 2,6-Me2C6H3 (Xyl), R′ = Tol, 3a; R = Xyl, R′ = Ph, 3b; R = Xyl, R′=Bun, 3c; R = Xyl, R′=SiMe3, 3d; R = Me, R′ = Tol, 3e; R = Me, R′ = Ph, 3f) are obtained in two steps by addition of R′CCLi (R′ = Tol, Ph, Bun, SiMe3) to the carbonyl aminocarbyne complexes [Fe2{μ-CN(Me)(R)}(μ-CO)(CO)2(Cp)2]+ (R = Xyl, 1a; Me, 1b), followed by methylation of the resulting alkynyl acyl compounds [Fe2{μ-CN(Me)(R)}(μ-CO)(CO){C(O)CCR′}(Cp)2] (R = Xyl, R′ = Tol, 2a; R = Xyl, R′ = Ph, 2b; R = Xyl, R′ = Bun, 2c; R = Xyl, R′ = SiMe3, 2d; R = Me, R′ = Tol, 2e; R = Me, R′ = Ph, 2f). Complexes 3 react with secondary amines (i.e., Me2NH, C5H10NH) to give the 4-amino-1-metalla-1,3-dienes [Fe2{μ-CN(Me)(R)}(μ-CO)(CO){C(OMe)CHC(R′)(NMe2)}(Cp)2]+ (R = Xyl, R′ = Tol, 4a; R = Xyl, R′ = Ph, 4b; R = Me, R′ = Ph, 4c) and [Fe2{μ-CN(Me)(Xyl)}(μ-CO)(CO){C(OMe)CHC(Tol)(NC5H10)}(Cp)2]+, 5. The addition occurs stereo-selectively affording only the E-configured products. Analogously, addition of primary amines R′NH2 (R′ = Ph, Et, Pri) affords the 4-(NH-amino)-1-metalla-1,3-diene complexes [Fe2{μ-CN(Me)(Xyl)}(μ-CO)(CO){C(OMe)CHC(R)(NHR′)}(Cp)2]+ (R = Ph, 6a; Et, 6b; Pri, 6c). In the case of 6a, only the E isomer is formed, whereas a mixture of the E and Z isomers is present in the case of 6b,c, with prevalence of the latter. Moreover, the two isomeric forms exist under dynamic equilibrium conditions, as shown by VT NMR studies. Complexes 6 are deprotonated by strong bases (e.g., NaH) resulting in the formation of the neutral vinyl imine complexes [Fe2{μ-CN(Me)(Xyl)}(μ-CO)(CO){C(OMe)CHC(NR)(Tol)}(Cp)2] (R = Ph, 7a; Et, 7b; Pri, 7c); the reaction can be reverted by addition of strong acids. X-ray crystal structures have been determined for 3a[CF3SO3] · Et2O, 4c[CF3SO3], 6a[BF4] · CH2Cl2, 6c[CF3SO3] · 0.5Et2O and 7a · CH2Cl2.  相似文献   

13.
The dinuclear nickel(II) complex [Ni2L(Cl)]+ (1), where (L)2− represents a 24-membered binucleating hexamine-dithiophenolate ligand, reacts readily with primary and secondary amines RR′NH in the presence of CO2 (1 bar) to give dinuclear monoalkyl- and dialkylcarbamate complexes [Ni2L(O2CNRR′)]+ (R = H, R′ = CH2Ph (2), R = H, R′ = n-Bu (3), R = H, R′ = n-Oct (4), R = H, R′ = CH2CH2OH (5), R = R′ = Et (6), and R = R′ = CH2CH2OH (7)). Complexes 2-7 can also be prepared by the reaction of 1 with CO2(air)/amine. The carbamate complexes are hydrolyzed in methanolic solution to give the known alkylcarbonate complex [Ni2L(O2COMe)]+ (8). These conversions are less rapid than the transesterification reactions of 8, due to a less electron-demanding carboxyl C(carbamate) atom. All new complexes were either isolated as perchlorate or tetraphenylborate salts and fully characterized by elemental analysis, UV/Vis, and IR spectroscopy. The structures of 2[BPh4] and 7[BPh4] have also been determined by X-ray crystallography. They confirm the presence of μ1,3-bridging alkylcarbamate units in the products.  相似文献   

14.
15.
The complexes [Cu2(ox)(phen)2(H2O)2](NO3)2 (1), [Cu2(sq)(pmdien)2(H2O)2](ClO4)2 (2) and {[Cu3(pdc)3(4,4′-bipy)1.5(H2O)2.25] · 2.5(H2O)}n (3) [phen = 1,10-phenanthroline; pmdien = N,N,N′,N′,N″-pentamethyldiethylenetriamine; 4,4′-bipy = 4,4′-bipyridine; ox = oxalate dianion; sq = squarate dianion and pdc = pyridine 2,6-dicarboxylate] have been synthesized and characterized by X-ray single crystal structure determination, low temperature magnetic measurement and thermal study. Structure determination reveals that 1 and 2 are dinuclear copper(II) complexes bridged by oxalate and squarate dianions, respectively, while 3 is a hexanuclear species formed by three Cu(pdc)(H2O)-(4,4′-bipy)-Cu(pdc)(H2O) fragments, connected through long Cu-O(pdc) bonds in a centrosymmetric arrangement. In complex 1 H-bonds occurring between the coordinated water molecules and lattice nitrate anions result in eight-membered ring clusters with the concomitant formation of 1D supramolecular chain. The adjacent chains undergo π-π stacking forming a 2D architecture. In the crystal of 3 an extensive H-bonding scheme gives rise to a 3D supramolecular network. Low temperature magnetic study shows a strong antiferromagnetic coupling in 1 (J = −288 ± 2 cm−1, g = 2.21 ± 0.01, R = 1.2 × 10−6); and a very weak interaction in 2 and 3, the best-fit parameters being: J = −0.21 cm−1, g = 2.12 ± 0.01, R = 1.1 × 10−6 (2) and J = −1.34 cm−1 ± 0.1, g = 2.14 ± 0.01, R = 1.2 × 10−6 (3) (R defines as .  相似文献   

16.
Several new mononuclear and dinuclear ruthenium(II) complexes - incorporating 2,2′:6′,2″-terpyridine and acetylacetonate as ancillary ligands and phenylcyanamide derivative ligands - of the type [Ru(tpy)(acac)(L)] and [{Ru(tpy)(acac)}2(μ-L′)] (where tpy = 2,2′:6′,2″-terpyridine, acac = acetylacetonate, L = hmbpcyd = 4-(3-hydroxy-3-methylbutynyl)phenylcyanamide anion (2) and epcyd = 4-ethynylphenylcyanamide anion (3) and L′ = bcpda = bis(4-cyanamidophenyl)diacetylene dianion (4) and bcpea = 9,10-bis(4-cyanamidophenylethynyl)anthracene dianion (5)) were synthesized in a stepwise manner starting from [Ru(tpy)(acac)(Ipcyd)] (1), where Ipcyd = 4-iodophenylcyanamide anion. Tetraphenylarsonium salts of the phenylcyanamide derivative ligands were also prepared. The four complexes have been characterized by UV-Vis, IR, ES-MS, electrochemistry and 1H NMR. Mononuclear complexes 2 and 3 were further characterized by 13C NMR. The single crystal X-ray structure of 2 was determined, it crystallized with one molecule of water with empirical formula of C32H31N5O5Ru, in a monoclinic crystal system and space group of P21/n with a = 17.642(5) Å, b = 9.634(2) Å, c = 20.063(7) Å, β = 92.65(3)°, V = 3406(2) Å3 and Z = 4. The structure was refined to a final R factor of 0.040. The Ru(III/II) couple of 1-3 appeared around 0.34 V versus the saturated calomel electrode in dimethylformamide and at a slightly higher potential, around 0.36-0.37 V for 4 and 5. Spectroelectrochemical studies were also performed for 4 and 5, no intervalence transition was observed despite all attempts.  相似文献   

17.
An unnatural base pair between 7-(2-thienyl)-imidazo[4,5-b]pyridine (Ds) and pyrrole-2-carbaldehyde (Pa) could expand the genetic alphabet and allow the incorporation of non-standard amino acids into proteins at defined positions. For this purpose, we synthesized tRNAs bearing Pa at the anticodon and tested non-standard amino acid phosphoserine aminoacylation by the wild-type and various engineered phosphoseryl-tRNA synthetases (SepRSs). The D418N D420N T423V triple mutant of SepRS efficiently charged phosphoserine to tRNA containing the PaUA anticodon with a Km = 47.1 μM and a kcat = 0.151 s−1, which are comparable to the values of the wild-type SepRS for its cognate substrate, tRNACys with the GCA anticodon (26.9 μM and 0.111 s−1). The triple mutant SepRS and the tRNA with the PaUA anticodon represent a specific pair for the site-specific incorporation of phosphoserine into proteins in response to the UADs codon within mRNA.  相似文献   

18.
Complexes cis,trans-Fe(CO)2(PMe3)2RR′ (R = CH3, R′ = Ph (2); R = CH3, R′ = CHCH2 (3); R = CHCH2, R′ = Ph (4); R = R′ = CHCH2 (5); R = R′ = CH3 (6)) were prepared by reaction of cis,trans-Fe(CO)2(PMe3)2RCl (1) with organolithium reagents LiR′. All complexes were characterized in solution by IR and 1H, 31P and, in a few cases, 13C NMR mono- and bi-dimensional spectroscopies. Complexes 5 and 6 were structurally characterized by X-ray diffractometric methods. In solution complexes 2, 3 and 4 undergo slowly coupling of the σ-hydrocarbyl substituents leading to Fe(CO)3(PMe3)2 and other decomposition products. Complex 6 was very stable in solution in the absence of nucleophiles and in the solid state. Complex 5 transformed through intramolecular coupling of the vinyl groups into Fe(CO)(PMe3)24-butadiene) (7), which was characterized in solution by IR and NMR spectroscopies.  相似文献   

19.
Hyper-pigmentation of the skin is a common problem that is prevalent in middle aged and elderly people. It is caused by over production of melanin. Tyrosinase is known to be the key enzyme in melanin production. Ethanolic extract of Greyia flanaganii leaves showed significant (P < 0.05) antityrosinase activity exhibiting the IC50 of 32.62 μg/ml. The total extract was further investigated for its toxicity and effect on melanin production by melanocytes cells, and showed significant inhibition (P < 0.05) (20%) of melanin production at 6.25 μg/ml and low levels of cytotoxicity (IC50 < 400 μg/ml). The amount of antioxidants necessary to decrease the initial DPPH absorbance by 50% (EC50) by the total ethanolic extract was found to be 22.01 μg/ml. The effect of G. flanaganii against acne causing bacteria, Propionibacterium acnes, was investigated using microdilution assay. The MIC of the extract of G. flanaganii was found to be 250 μg/ml. Bioassay-guided fractionation led to the isolation of (3S)-4-hydroxyphenethyl 3-hydroxy-5-phenylpentanoate (1), 2′,4′,6′-trihydroxydihydrochalcone (2), 2′,6′,4-trihydroxy-4′-methoxydihydrochalcone (3), 2′,6′-dihydroxy-4′-methoxydihydrochalcone (4), 5,7-dihydroxyflavanone [(2S)-pinocembrin] (5), 2′,6′-dihydroxy-4′,4-dimethoxy dihydrochalcone (6) and (2R,3R)-3,5,7-trihydroxy-3-O-acetylflavanone (7). The isolated compounds were tested for their antioxidant, cytotoxicity, tyrosinase inhibition and antibacterial activities. Compound 2 exhibited significant (P < 0.05) antityrosinase activity exhibiting the IC50 of 69.15 μM. The isolated compounds showed low toxicity of the cells with reduction of melanin content of the cells. All compounds tested showed good radical scavenging activity. These data indicates that G. flanaganii extract and its isolated phenolic constituents could be possible skin lightening agents.  相似文献   

20.
The synthesis and characterization of several complexes of the composition [{M(terpy)}n(L)](ClO4)m (M = Pt, Pd; L = 1-methylimidazole, 1-methyltetrazole, 1-methyltetrazolate; terpy = 2,2′:6′,2″-terpyridine; n = 1, 2; m = 1, 2, 3) is reported and their applicability in terms of a metal-mediated base pair investigated. Reaction of [M(terpy)(H2O)]2+ with 1-methylimidazole leads to [M(terpy)(1-methylimidazole)](ClO4)2 (1: M = Pt; 2: M = Pd). The analogous reaction of [Pt(terpy)(H2O)]2+ with 1-methyltetrazole leads to the organometallic compound [Pt(terpy)(1-methyltetrazolate)]ClO4 (3) in which the aromatic tetrazole proton has been substituted by the platinum moiety. For both platinum(II) and palladium(II), doubly metalated complexes [{M(terpy)}2(1-methyltetrazolate)](ClO4)3 (4: M = Pt; 5: M = Pd) can also be obtained depending on the reaction conditions. In the latter two compounds, the [M(terpy)]2+ moieties are coordinated via C5 and N4. X-ray crystal structures of 1, 2, and 3 are reported. In addition, DFT calculations have been carried out to determine the energy difference between fully planar [Pd(mterpy)(L)]2+ complexes Ip-IVp (mterpy = 4′-methyl-2,2′:6′,2″-terpyridine; L = 1-methylimidazole-N3 (I), 1-methyl-1,2,4-triazole-N4 (II), 1-methyltetrazole-N3 (III), or 3-methylpyridine-N1 (IV)) and the respective geometry-optimized structures Io-IVo. Whereas this energy difference is larger than 70 kJ mol−1 for compounds I, II, and IV, it amounts to only 0.8 kJ mol−1 for the tetrazole-containing complex III, which is stabilized by two intramolecular C-H?N hydrogen bonds. Of all complexes under investigation, only the terpyridine-metal ion-tetrazole system with N3-coordinated tetrazole appears to be suited for an application in terms of a metal-mediated base pair in a metal-modified oligonucleotide.  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