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 15 毫秒
1.
The reaction of FeCl(2) or FeCl(3) with sodium ethanethiolate (SEt) in N-methylformamide (NMF) has been reevaluated to rectify a previous Fe(II) oxidation artifact. On titrating Fe(II) with EtS(-) concentrations up to 12 mol Eq, new features in the UV/vis spectrum (epsilon(344)=(3.1+/-0.2)x10(3) M(-1) cm(-1); epsilon(486)=(4.5+/-0.1)x10(2) M(-1) cm(-1)) indicated that the first observable step was the formation of a single complex different from the known tetrahedral tetrathiolate, [Fe(SEt)(4)](2-) . As the EtS(-) concentration increased past 12.5 mol Eq the UV/vis spectrum gradually transformed to that of [Fe(SEt)(4)](2-) (lambda(max)=314 nm). A Hill-formalism fit to the titration data of the initially formed complex indicated cooperative ligation by three ethanethiolate ions, with K(coop)=(1.7+/-0.1)x10(3) M(-3) and Hill "n"=2.4+/-0.1 (r=0.997). The 3:1 EtS(-)-Fe(II) complex is proposed to be [Fe(2)(SEt)(6)](2-). Titration of Fe(III) with EtS(-) showed direct cooperative formation of [Fe(SEt)(4)](-) [epsilon(340)=(3.4+/-0.5)x10(3) M(-1) cm(-1)] with a Hill-formalism K(coop)=(4.3+/-0.1)x10(2) M(-4) and a Hill coefficient "n"=3.7+/-0.2 (r=0.996). Further ligation past [Fe(SEt)(4)](-) was observed at EtS(-) concentrations above 35 mol Eq. The Fe(III) Hill constants are at variance with our previous report. However, the UV/vis spectrum of Fe(III) in NMF solution was found to change systematically over time, consistent with a slow progressive deprotonation of [Fe(nmf)](3+). The observed time-to-time differences in the equilibrium chemistry of Fe(III) with ethanethiolate in NMF thus reflect variation in the microscopic solution composition of FeCl(3) in alkaline NMF solvent. These results are related to the chemistry of nitrogenase FeMo cofactor in alkaline NMF solution.  相似文献   

2.
Cu in blood has been believed to transport into cell via albumin and some amino acids. To shed light on the Cu transport process we studied the reaction of the Cu(II)-peptide with the amino acid by absorption and CD spectra. Albumin mimic peptides GlyGly-L-HisGly (GGHG) and penta-Gly(G5) formed stable 4N coordinated Cu(II) complexes, but in the reaction with histidine (His) and penicillamine (Pes) the ternary Cu(II) complex formations were observed different by the kinetic study. Cu(II)-G5 complexes reacted with Pes to form the ternary complex Cu(H(-1)G5)(Pes(-)) which was subsequently transformed to the binary complex Cu(Pes(-))(2). In the system with GGHG the Cu(II) was also transported from GGHG to Pes, but the ternary Cu(H(-1)GGHG)(Pes(-)) complex as the intermediate was detected a trace. The ternary complex would be spontaneously transformed to Cu(Pes(-))(2) upon forming, because the rate constant of the ternary complex formation k(1+)= approximately 2M(-1)s(-1) was less than k(2+)= approximately 5 x 10(2)M(-1)s(-1) for the Cu(Pes(-))(2) formation at physiological pH. In the Cu(II)-GGHG-His system the ternary Cu(H(-1)GGHG)(His) complex was also hardly identified because the formation constant K(1) and k(1+) were very small and the equilibrium existed between Cu(H(-2)GGHG) and Cu(His)(2) and its overall equilibrium constant beta(2) for Cu(His)(2) was very small to be 1.00+/-0.05 M(-1) at pH 9.0. These results indicated that the ternary complex is formed in the Cu transport process from the albumin to the amino acid, but His imidazole nitrogen in the fourth-binding site of Cu(II) strongly resists the replacement by the incoming ligand.  相似文献   

3.
The enzyme L-asparaginase (ASNASE), which hydrolyzes L-asparagine (L-Asn) to ammonia and L-aspartic acid (L-Asp), is commonly used for remission induction in acute lymphoblastic leukemia. To correlate ASNASE activity with L-Asn reduction in human serum, sensitive methods for the determination of ASNASE activity are required. Using L-aspartic beta-hydroxamate (AHA) as substrate we developed a sensitive plate reader-based method for the quantification of ASNASE derived from Escherichia coli and Erwinia chrysanthemi and of pegylated E. coli ASNASE in human serum. ASNASE hydrolyzed AHA to L-Asp and hydroxylamine, which was determined at 710 nm after condensation with 8-hydroxyquinoline and oxidation to indooxine. Measuring the indooxine formation allowed the detection of 2 x 10(-5)U ASNASE in 20 microl serum. Linearity was observed within 2.5-75 and 75-1,250 U/L with coefficients of correlation of r(2)>0.99. The coefficients of variation for intra- and interday variability for the three different ASNASE enzymes were 1.98 to 8.77 and 1.73 to 11.0%. The overall recovery was 101+/-9.92%. The coefficient of correlation for dilution linearity was determined as r(2)=0.986 for dilutions up to 1:20. This method combined with sensitive methods for the quantification of L-Asn will allow bioequivalence studies and individualized therapeutic drug monitoring of different ASNASE preparations.  相似文献   

4.
The enzymatic characterization of GDP-d-mannose 3',5'-epimerase (GME), a key enzyme in the biosynthesis of vitamin C in plants is described. The GME gene (Genbank Accession No. AB193582) in rice was cloned, and expressed as a fusion protein in Escherichia coli. Reaction products from GDP-d-mannose, as produced by GME catalysis, were separated by recycling HPLC on an ODS column, and were determined to be GDP-l-galactose and GDP-l-gulose, based on their NMR spectra and sugar analysis. The reaction catalyzed by GME was inhibited by GDP, and was strongly accelerated by NAD(+) in contrast to the case of GME from Arabidopsis thaliana. This difference in the effect of NAD(+) on GME activity can be attributed to the NAD binding domain which is conserved in the rice gene, but not in the Arabidopsis thaliana gene. The apparent K(m) and k(cat) were determined to be 1.20x10(-5)M and 0.127s(-1), respectively, in the presence of 20microM NAD(+). The fractions of GDP-d-mannose, GDP-l-galactose and GDP-l-gulose, at equilibrium, were approximately 0.75, 0.20 and 0.05, respectively.  相似文献   

5.
During iron acquisition by the cell, complete homodimeric transferrin receptor 1 in an unknown state (R1) binds iron-loaded human serum apotransferrin in an unknown state (T) and allows its internalization in the cytoplasm. T also forms complexes with metals other than iron. Are these metals incorporated by the iron acquisition pathway and how can other proteins interact with R1? We report here a four-step mechanism for cobalt(III) transfer from CoNtaCO(3)(2-) to T and analyze the interaction of cobalt-loaded transferrin with R1. The first step in cobalt uptake by T is a fast transfer of Co(3+) and CO(3)(2-) from CoNtaCO(3)(2-) to the metal-binding site in the C-lobe of T: direct rate constant, k(1)=(1.1+/-0.1) x 10(6) M(-1) s(-1); reverse rate constant, k(-1)=(1.9+/-0.6) x 10(6) M(-1) s(-1); and equilibrium constant, K=1.7+/-0.7. This step is followed by a proton-assisted conformational change of the C-lobe: direct rate constant, k(2)=(3+/-0.3) x 10(6) M(-1) s(-1); reverse rate constant, k(-2)=(1.6+/-0.3) x 10(-2) s(-1); and equilibrium constant, K(2a)=5.3+/-1.5 nM. The two final steps are slow changes in the conformation of the protein (0.5 h and 72 h), which allow it to achieve its final thermodynamic state and also to acquire second cobalt. The cobalt-saturated transferrin in an unknown state (TCo(2)) interacts with R1 in two different steps. The first is an ultra-fast interaction of the C-lobe of TCo(2) with the helical domain of R1: direct rate constant, k(3)=(4.4+/-0.6)x10(10) M(-1) s(-1); reverse rate constant, k(-3)=(3.6+/-0.6) x 10(4) s(-1); and dissociation constant, K(1d)=0.82+/-0.25 muM. The second is a very slow interaction of the N-lobe of TCo(2) with the protease-like domain of R1. This increases the stability of the protein-protein adduct by 30-fold with an average overall dissociation constant K(d)=25+/-10 nM. The main trigger in the R1-mediated iron acquisition is the ultra-fast interaction of the metal-loaded C-lobe of T with R1. This step is much faster than endocytosis, which in turn is much faster than the interaction of the N-lobe of T with the protease-like domain. This can explain why other metal-loaded transferrins or a protein such as HFE-with a lower affinity for R1 than iron-saturated transferrin but with, however, similar or higher affinities for the helical domain than the C-lobe-competes with iron-saturated transferrin in an unknown state towards interaction with R1.  相似文献   

6.
Proton NMR spectroscopy was applied to study the reactions of the dipeptides glycyl-glycine (Gly-Gly) and glycyl-l-alanine (Gly-l-Ala) with hydrogen tetrachloridoaurate(III) (H[AuCl4]). All reactions were performed at pH 2.0 and 3.0 and at 40 °C. The final products in these reactions were [Au(Gly-Gly-κ3NG1,NG2,OG2)Cl] and [Au(Gly-l-Ala-κ3NG,NA,OA)Cl] complexes. Tridentate coordination of the corresponding dipeptides and square-planar geometry of these Au(III) complexes was confirmed by NMR (1H and 13C) spectroscopy. This study showed that at pH < 3.0 the Au(III) ion was able to deprotonate the amide nitrogen atom. However this displacement reaction was very slow and the total concentration of the corresponding Au(III)-peptide complex formed after 5 days was less than 60% for the Gly-l-Ala or 70% for the Gly-Gly dipeptide. The kinetic data of the reactions between the Gly-Gly and Gly-l-Ala dipeptides and [AuCl4] were compared with those for the histidine-containing Gly-l-His dipeptide. The differences in the reactivity of these three dipeptides with the Au(III) ion are discussed.  相似文献   

7.
The oxidation of D-lactobionic acid by Cr(VI) yields the 2-ketoaldobionic acid and Cr(3+) as final products when a 20-times or higher excess of the aldobionic acid over Cr(VI) is used. The redox reaction takes place through a complex multistep mechanism, which involves the formation of intermediate Cr(IV) and Cr(V) species. Cr(IV) reacts with lactobionic acid much faster than Cr(V) and Cr(VI) do, and cannot be directly detected. However, the formation of CrO(2)(2+), observed by the first time for an acid saccharide/Cr(VI) system, provides indirect evidence for the intermediacy of Cr(IV) in the reaction path. Cr(VI) and the intermediate Cr(V) react with lactobionic acid at comparable rates, being the complete rate laws for the Cr(VI) and Cr(V) consumption expressed by: -d[Cr(VI)]/dt=[k(I)+k(II)[H(+)]][lactobionicacid][Cr(VI)], where k(I)=(4.1+/-0.1) x 10(-3) M(-1) s(-1) and k(II)=(2.1+/-0.1) x 10(-2) M(-2) s(-1); and -d[Cr(V)]/dt=[k(III)[H(+)]+(k(IV)+k(V)[H(+)])[lactobionicacid]] [Cr(V)], where k(III)=(1.8+/-0.1) x 10(-3) M(-1) s(-1), k(IV)=(1.1+/-0.1) x 10(-2) M(-1) s(-1) and k(V)=(1.0+/-0.1) x 10(-2) M(-2) s(-1), at 33 degrees C. The Electron Paramagnetic Resonance (EPR) spectra show that five-co-ordinate oxo-Cr(V) bischelates are formed at pH 1-5 with the aldobionic acid bound to Cr(V) through the alpha-hydroxyacid group.  相似文献   

8.
A water-insoluble alpha-(1-->3)-D-glucan (A) from Lentinus edodes was fractionated into 13 fractions in dimethyl sulfoxide containing 0.25 M lithium chloride (0.25 M LiCl-Me(2)SO). Five fractions were treated with sulfur trioxide-pyridine complex at 25 degrees C to synthesize water-soluble sulfated derivatives (S-A). The weight-average molecular weights, M(w), and intrinsic viscosities [eta], of the samples A and S-A were determined by multi-angler laser light scattering (MALLS), and viscosity. The M(w) dependence of [eta] and of the radius of gyration (z)(1/2), was found to be represented approximately by [eta]=4.9 x 10(-2) M(w)(0.67) (cm(3) g(-1)), and (z)(1/2)=4.8 x 10(-2) M(w)(0.54) (nm) for the alpha-glucan in 0.25 M LiCl-Me(2)SO in the M(w) range from 7.24 x 10(4) to 4.21 x 10(5), and by [eta]=6.8 x 10(-4) M(w) 1.06 (cm(3) g(-1)), and (z)(1/2)=9.4 x 10(-4) M(w)(0.92) (nm) for the sulfated alpha-glucan in aqueous 0.5 M NaCl in the M(w) range from 5.92 x 10(4) to 1.42 x 10(5) at 25 degrees C. The results indicate that the alpha-(1-->3)-D-glucan exists as a flexible chain in 0.25 M LiCl-Me(2)SO, and its sulfated derivative in 0.5 M NaCl aqueous has stiffer chains than the original. (13)C NMR indicated that intramolecular hydrogen bonding occurred in the sulfated alpha-glucan, causing the observed chain stiffness.  相似文献   

9.
Glycosidation of 2,5-anhydro-1,6-di-O-benzoyl-D-mannitol with methyl(2,3,4-tri-O-acetyl-alpha-d-glucopyranosyl-1-O-trichloroacetimidate)uronate in the presence of trimethylsilyl triflate afforded the corresponding 3-O-beta-glycoside, which after deprotection was converted into its hexa-O-sulfate with DMF x SO3 to give after treatment with sodium acetate and subsequent saponification of the methyl ester with sodium hydroxide the hepta sodium salt of 2,5-anhydro-3-O-(beta-d-glucopyranosyl uronate)-D-mannitol hexa-O-sulfate. Glycosidation of the same acceptor with the alpha-thiophenylglycoside of methyl 2,4-di-O-acetyl-3-O-benzyl-L-idopyranosyl uronate in the presence of NIS/TfOH afforded the corresponding 3-O-alpha-glycoside in very low yield, therefore the alpha-thiophenylglycoside of 2-O-acetyl-2,4-O-benzylidene-3-O-benzyl-L-idopyranose was used as donor. The terminal hydroxymethyl group of the obtained disaccharide was subsequently oxidised with NaOCl/TEMPO and the obtained iduronic acid derivative was converted into the hepta sodium salt of 2,5-anhydro-3-O-(-alpha-L-idopyranosyluronate)-D-mannitol hexa-O-sulfonate with DMF x SO3 and subsequent treatment with sodium acetate.  相似文献   

10.
11.
It was found by 1H and 13C NMR spectroscopy that the Schiff base, 2-deoxy-2-(2-hydroxybenzaldimino)-D-glucopyranose exhibits enol-imine-keto-amine and anomeric equilibria in methanolic, and in dimethyl sulfoxide solutions. The reaction of the Schiff base with nickel acetate gave the bidentate, mononuclear Ni(II) complex that was characterized by spectroscopic methods and by cyclic voltammetry. The coordination of the Schiff base to the metal is through the enol-imine tautomeric form, and the anomeric equilibrium remains in dimethyl sulfoxide solutions. This complex was also obtained by reaction of D-glucosamine with Ni(II) salicylaldehydate. The same reaction was employed for the synthesis of bis-N-[2-deoxy-D-galactopyranosyl-2-(2-hydroxybenzaldiminate)]Ni(II). The small paramagnetic shifts of the 1H NMR resonances of the complexes suggest that paramagnetic species are present in low proportions.  相似文献   

12.
The reaction of a racemic mixture of (2R,2'S)- and (2S,2'R)-N-(p-tolylsulfonyl)-2-pyrrolidinyl-2-propanol, prepared from (S)-proline, with 2,3,4-tri-O-acetyl-alpha-L-fucopyranosyl trichloroacetimidate led to both diastereoisomers of the title compound after O-deacetylation.  相似文献   

13.
L-Arabinose isomerase (AI) catalyzes the isomerization of L-arabinose to L-ribulose. It can also convert d-galactose to d-tagatose at elevated temperatures in the presence of divalent metal ions. The araA genes, encoding AI, from the mesophilic bacterium Bacillus halodurans and the thermophilic Geobacillus stearothermophilus were cloned and overexpressed in Escherichia coli, and the recombinant enzymes were purified to homogeneity. The purified enzymes are homotetramers with a molecular mass of 232 kDa and close amino acid sequence identity (67%). However, they exhibit quite different temperature dependence and metal requirements. B. halodurans AI has maximal activity at 50 degrees C under the assay conditions used and is not dependent on divalent metal ions. Its apparent K(m) values are 36 mM for L-arabinose and 167 mM for d-galactose, and the catalytic efficiencies (k(cat)/K(m)) of the enzyme were 51.4 mM(-1)min(-1) (L-arabinose) and 0.4 mM(-1)min(-1) (d-galactose). Unlike B. halodurans AI, G. stearothermophilus AI has maximal activity at 65-70 degrees C, and is strongly activated by Mn(2+). It also has a much higher catalytic efficiency of 4.3 mM(-1)min(-1) for d-galactose and 32.5 mM(-1)min(-1)for L-arabinose, with apparent K(m) values of 117 and 63 mM, respectively. Irreversible thermal denaturation experiments using circular dichroism (CD) spectroscopy showed that the apparent melting temperature of B. halodurans AI (T(m)=65-67 degrees C) was unaffected by the presence of metal ions, whereas EDTA-treated G. stearothermophilus AI had a lower T(m) (72 degrees C) than the holoenzyme (78 degrees C). CD studies of both enzymes demonstrated that metal-mediated significant conformational changes were found in holo G. stearothermophilus AI, and there is an active tertiary structure for G. stearothermophilus AI at elevated temperatures for its catalytic activity. This is in marked contrast to the mesophilic B. halodurans AI where cofactor coordination is not necessary for proper protein folding. The metal dependence of G. stearothermophilus AI seems to be correlated with their catalytic and structural functions. We therefore propose that the metal ion requirement of the thermophilic G. stearothermophilus AI reflects the need to adopt the correct substrate-binding conformation and the structural stability at elevated temperatures.  相似文献   

14.
The [Ru(II)(Hedta)NO(+)] complex is a diamagnetic species crystallizing in a distorted octahedral geometry, with the Ru-N(O) length 1.756(4) A and the RuNO angle 172.3(4) degrees . The complex contains one protonated carboxylate (pK(a)=2.7+/-0.1). The [Ru(II)(Hedta)NO(+)] complex undergoes a nitrosyl-centered one-electron reduction (chemical or electrochemical), with E(NO+/NO)=-0.31 V vs SCE (I=0.2 M, pH 1), yielding [Ru(II)(Hedta)NO](-), which aquates slowly: k(-NO)=2.1+/-0.4x10(-3) s(-1) (pH 1.0, I=0.2 M, CF(3)COOH/NaCF(3)COO, 25 degrees C). At pHs>12, the predominant species, [Ru(II)(edta)NO](-), reacts according to [Ru(II)(edta)NO](-)+2OH(-)-->[Ru(II)(edta)NO(2)](3-), with K(eq)=1.0+/-0.4 x 10(3) M(-2) (I=1.0 M, NaCl; T=25.0+/-0.1 degrees C). The rate-law is first order in each of the reactants for most reaction conditions, with k(OH(-))=4.35+/-0.02 M(-1)s(-1) (25.0 degrees C), assignable mechanistically to the elementary step comprising the attack of one OH(-) on [Ru(II)(edta)NO](-), with subsequent fast deprotonation of the [Ru(II)(edta)NO(2)H](2-) intermediate. The activation parameters were DeltaH(#)=60+/-1 kJ/mol, DeltaS(#)=-31+/-3 J/Kmol, consistent with a nucleophilic addition process between likely charged ions. In the toxicity up-and-down tests performed with Swiss mice, no death was observed in all the doses administered (3-9.08 x 10(-5) mol/kg). The biodistribution tests performed with Wistar male rats showed metal in the liver, kidney, urine and plasma. Eight hours after the injection no metal was detected in the samples. The vasodilator effect of [Ru(II)(edta)NO](-) was studied in aortic rings without endothelium, and was compared with sodium nitroprusside (SNP). The times of maximal effects of [Ru(II)(edta)NO](-) and SNP were 2 h and 12 min, respectively, suggesting that [Ru(II)(edta)NO](-) releases NO slowly to the medium in comparison with SNP.  相似文献   

15.
The binding of Ru(phen)(2)dppz(2+) (dppz=dipyrido[3,2-a:2',3'-c]phenazine) to DNA was investigated at pH 7.0 and 25 degrees C using stopped-flow and spectrophotometric methods. Equilibrium measurements show that two modes of binding, whose characteristics depend on the polymer to dye ratio (C(P)/C(D)), are operative. The binding mode occurring for values of C(P)/C(D) higher than 3 exhibits positive cooperativity, which is confirmed by kinetic experiments. The reaction parameters are K=2 x 10(3)M(-1), omega=550, n=1, k(r)=(1.9+/-0.5) x 10(7)M(-1)s(-1) and k(d)=(9.5+/-2.5)x10(3)s(-1) at I=0.012 M. The results are discussed in terms of prevailing surface interaction with DNA grooves accompanied by partial intercalation of the dppz residue. The other binding mode becomes operative for C(P)/C(D)<3 and the equilibria analysis shows this is an ordinary intercalation mode (K=1.3 x 10(6) M(-1), n=1.5 at I=0.012 M and K=2 x 10(5) M(-1), n=1.2 at I=0.21 M). Similar behaviour is displayed by double-stranded poly(A).  相似文献   

16.
In previous studies [FEBS Lett. 434 (1998) 231, Arch. Biochem. Biophys. 404 (2002) 92], we demonstrated for the first time that D-aspartate (D-Asp) is synthesized in cultured mammalian cell lines, such as pheochromocytoma 12 (PC12) and its subclone, MPT1. Our current focus is analysis of the dynamics of D-Asp homeostasis in these cells. In this communication, we show that L-glutamate (Glu) and L-Glu transporter substrates in the extracellular space regulate the homeostasis of endogenous D-Asp in MPT1 cells. D-Asp is apparently in dynamic homeostasis, whereby endogenous D-Asp is constantly released into the extracellular space by an undefined mechanism, and continuously and intensively taken up into cells by an L-Glu transporter. Under these conditions, L-Glu and its transporter substrates in the medium may competitively inhibit the uptake of D-Asp via the transporter, resulting in accumulation of the amino acid in the extracellular space. We additionally demonstrate that DL-TBOA, a well-established L-Glu transporter inhibitor, is taken up by the transporter during long time intervals, but not on a short time-scale.  相似文献   

17.
An efficient and practical route for the large-scale synthesis of 2-deoxy-L-erythro-pentose (2-deoxy-L-ribose) starting from L-arabinose was developed using Barton-type free-radical deoxygenation reaction as a key step. The radical precursor, a phenoxythiocarbonyl ester, was prepared in situ, and the most efficient deoxygenation was achieved by slow addition of tributyltin hydride to the reaction mixture.  相似文献   

18.
The transport of l-leucine, l-phenylalanine and l-alanine by the perfused lactating rat mammary gland has been examined using a rapid, paired-tracer dilution technique. The clearances of all three amino acids by the mammary gland consisted of a rising phase followed by a rapid fall-off, respectively, reflecting influx and efflux of the radiotracers. The peak clearance of l-leucine was inhibited by BCH (65%) and d-leucine (58%) but not by l-proline. The inhibition of l-leucine clearance by BCH and d-leucine was not additive. l-leucine inhibited the peak clearance of radiolabelled l-leucine by 78%. BCH also inhibited the peak clearance of l-phenylalanine (66%) and l-alanine (33%) by the perfused mammary gland. Lactating rat mammary tissue was found to express both LAT1 and LAT2 mRNA. The results suggest that system L is situated in the basolateral aspect of the lactating rat mammary epithelium and thus probably plays a central role in neutral amino acid uptake from blood. The finding that l-alanine uptake by the gland was inhibited by BCH suggests that LAT2 may make a significant contribution to neutral amino acid uptake by the mammary epithelium.  相似文献   

19.
5'-[N-[(3S)-3-Amino-carboxypropyl]-N-methylamino]-5(')-deoxyadenosine (azaSAM), an analog of S-adenosyl-L-methionine (SAM), was used to study the cofactor-dependent reduction of the [4Fe-4S](2+) center in lysine 2,3-aminomutase to the +1 oxidation state. azaSAM has a tertiary nitrogen in place of the sulfonium center of SAM. The analog binds to lysine 2,3-aminomutase with K(d)s of 1.4+/-0.3 microM at pH 8.0 and 2.2+/-0.6 microM at pH 6.5. Reduction of the [4Fe-4S](2+) center in the presence of this analog gives a 10K [4Fe-4S](1+) electron paramagnetic resonance (EPR) signal similar to that seen with SAM or S-adenosyl-L-homocysteine (SAH). The pH dependence of cofactor-induced reduction was examined to determine whether ionization of the tertiary nitrogen (pK(a)=7.08) might affect reduction of the [4Fe-4S](2+) center. The results show similar behavior in azaSAM and SAH, demonstrating that ionization of the aza group in azaSAM does not account for pH dependence in cofactor-dependent reduction of the [4Fe-4S](2+) center. The signal shape of the low-temperature EPR signal for the [4Fe-4S](1+) center in the SAM-induced reduction displayed a pH dependence that was not observed in the azaSAM- or SAH-induced spectra. Unique features of the signal are at a maximum at the pH activity optimum of pH 8 and are diminished as the pH is lowered or raised. These features are also absent in the spectra at all pHs examined when reduction is induced by azaSAM or SAH.  相似文献   

20.
Ternary S-methyl-L-cysteine (SMe-l-cys) copper(II) complexes [Cu(SMe-L-cys)(B)(H(2)O)](X) (1-4), where the heterocyclic base B is 2,2'-bipyridine (bpy, 1), 1,10-phenanthroline (phen, 2), dipyridoquinoxaline (dpq, 3) and dipyridophenazine (dppz, 4), and X is ClO(4)(-) (1-3) or NO(3)(-) (4), are prepared and their DNA binding and cleavage properties studied. Complexes 2 and 4 are structurally characterized by X-ray crystallography. Both the crystal structures show distorted square-pyramidal (4+1) CuN(3)O(2) coordination geometry of the complexes in which the N,O-donor S-methyl-L-cysteine and N,N-donor heterocyclic base bind at the basal plane with a water molecule as the axial ligand. In addition, the dppz structure shows the presence of a 1D-chain formed due to covalent linkage of the carboxylate oxygen atom belonging to another molecule at the elongated axial site. The crystal structures show chemically significant non-covalent interactions like hydrogen bonding involving the axial aqua ligand and pi-pi interactions between dppz ligands. The complexes display a d-d band in the range of 605-654 nm in aqueous dimethylformamide (DMF) solution (9:1 v/v). The redox active complexes show quasireversible cyclic voltammetric response near 0.1 V in DMF assignable to the Cu(II)/Cu(I) couple. The complexes show good binding affinity to calf thymus (CT) DNA giving the order: 4 (dppz)>3 (dpq)>2 (phen)>1 (bpy). The intrinsic binding constants, obtained from UV-visible spectroscopic studies, are 1.3x10(4) and 2.15 x 10(4) M(-1) for 3 and 4, respectively. Control DNA cleavage experiments using pUC19 supercoiled (SC) DNA and minor groove binder distamycin suggest major groove binding propensity for the dppz complex, while the phen and dpq complexes bind at the minor groove of DNA. Complexes 2-4 show DNA cleavage activity in dark in the presence of a reducing agent 3-mercaptopropionic acid (MPA) via a mechanistic pathway involving formation of hydroxyl radical as the reactive species. The complexes also show efficient photo-induced DNA cleavage activity on irradiation with a monochromatic UV light of 365 nm in absence of any external reagent. The cleavage efficiency follows the order: 3>4>2. The complexes exhibit significant DNA cleavage activity on irradiation with visible light of 633 nm. Control experiments show inhibition of cleavage in presence of singlet oxygen quenchers like sodium azide, histidine and enhancement of cleavage in D(2)O, suggesting formation of singlet oxygen as a reactive species in a type-II process. The photosensitizing effect of the thiomethyl group of the amino acid is evidenced from the observation of significant DNA photocleavage activity of the phen complex 2 as the phen ligand itself is not a photosensitizer.  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