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 765 毫秒
1.
The reaction rate constants and transient spectra of 11 flavonoids and 4 phenolic acids reacting with e(aq)- at neutral pH were measured. Absorption bands of the transients of e(aq)- reacting with the above compounds all located at a wavelength shorter than 400 nm. The e(aq)- scavenging abilities were divided into three groups: (+)catechin ((1.2 +/-0.1) x 10(8) M(-1)s(-1)) < 4-chromanol ((4.4 +/- 0.4) x 10(8) M(-1)s(-1)) < genistein ((6.2+/-0.4) x 10(9) M (-1) s(-1) approximately genistin ((8 +/- 1) x 10(9) M(-1)s(-1)) approximately rutin ((7.6 +/- 0.4) x M(-1)s(-1) approximately caffeic acid ((8.3 +/- 0.5) x 10(9)M(-1)s(-1)) < transcinnamic acid((1.1 +/- 0.1) x 10(10) M(-1)s(-1)) approximately p-coumaric acid ((1.1 +/- 0.1) x 10(10) M(-1)s(-1) approximately 2,4,6-trihydroxylbenzoic acid((1.1 +/- 0.1) x 10(10) M(-1)s(-1)) approximately baicalein ((1.1 +/- 0.5) x 10(10) M(-1)s(-1)) approximately baicalin((1.3 + 0.1) X 10(10) M(-1)s(-1)) approximately naringenin ((1.2 +/- 0.1) x 10(10) M(-1)s(-1)) approximately naringin ((1.0 +/- 0.1) x 10(10) M(-1)s(-1)) approximately gossypin((1.2 +/- 0.1) x 10(10) M(-1)s(-1)) approximately quercetin((1.3 +/- 0.5) x 10(10) M(-1)s(-1)). These results suggested that C4 keto group is the active site for e(aq)- to attack on flavonoids and phenolic acids, whereas the o-dihydroxy structure in B ring, the C2,3 double bond, the C3-OH group, and glucosylation, which are key structures that influence the antioxidant activities of flavonoids and phenolic acids, have little effects on the e(aq)- scavenging activities.  相似文献   

2.
Pollock VV  Barber MJ 《Biochemistry》2001,40(5):1430-1440
Rhodobacter sphaeroides f. sp. denitrificans biotin sulfoxide reductase catalyzes the reduction of d-biotin d-sulfoxide (BSO) to biotin. Initial rate studies of the homogeneous recombinant enzyme, expressed in Escherichia coli, have demonstrated that the purified protein utilizes NADPH as a facile electron donor in the absence of any additional auxiliary proteins. We have previously shown [Pollock, V. V., and Barber, M. J. (1997) J. Biol. Chem. 272, 3355-3362] that, at pH 8 and in the presence of saturating concentrations of BSO, the enzyme exhibits, a marked preference for NADPH (k(cat,app) = 500 s(-1), K(m,app) = 269 microM, and k(cat,app)/K(m,app) = 1.86 x 10(6) M(-1) s(-1)) compared to NADH (k(cat,app) = 47 s(-1), K(m,app) = 394 microM, and k(cat,app)/K(m,app) = 1.19 x 10(5) M(-1) s(-1)). Production of biotin using NADPH as the electron donor was confirmed by both the disk biological assay and by reversed-phase HPLC analysis of the reaction products. The purified enzyme also utilized ferricyanide as an artificial electron acceptor, which effectively suppressed biotin sulfoxide reduction and biotin formation. Analysis of the enzyme isolated from tungsten-grown cells yielded decreased reduced methyl viologen:BSO reductase, NADPH:BSO reductase, and NADPH:FR activities, confirming that Mo is required for all activities. Kinetic analyses of substrate inhibition profiles revealed that the enzyme followed a Ping Pong Bi-Bi mechanism with both NADPH and BSO exhibiting double competitive substrate inhibition. Replots of the 1/v-axes intercepts of the parallel asymptotes obtained at several low concentrations of fixed substrate yielded a K(m) for BSO of 714 and 65 microM for NADPH. In contrast, utilizing NADH as an electron donor, the replots yielded a K(m) for BSO of 132 microM and 1.25 mM for NADH. Slope replots of data obtained at high concentrations of BSO yielded a K(i) for BSO of 6.10 mM and 900 microM for NADPH. Kinetic isotope studies utilizing stereospecifically deuterated NADPD indicated that BSO reductase uses specifically the 4R-hydrogen of the nicotinamide ring. Cyanide inhibited NADPH:BSO and NADPH:FR activities in a reversible manner while diethylpyrocarbonate treatment resulted in complete irreversible inactivation of the enzyme concomitant with molybdenum cofactor release, indicating that histidine residues are involved in cofactor-binding.  相似文献   

3.
Radiolabeling of biologically active molecules with the [(99m)Tc(CO)(3)](+) unit has been of primary interest in recent years. With this in mind, we herein report symmetric (L(1)) and asymmetric (L(2)-L(5)) pyrazolyl-containing chelators that have been evaluated in radiochemical reactions with the synthon [(99m)Tc(H(2)O)(3)(CO)(3)](+) (1a). These reactions yielded the radioactive building blocks [(99m)Tc(CO)(3)(k(3)-L)](+) (L = L(1)-L(5), 2a-6a), which were identified by RP-HPLC. The corresponding Re surrogates (2-6) allowed for macroscopic identification of the radiochemical conjugates. Complexes 2a-6a, with log P(o/w) values ranging from -2.35 to 0.87, were obtained in yields of > or =90% using ligand concentrations in the 10(-5-)10(-4) M range. Challenge studies with cysteine and histidine revealed high stability for all of these radioactive complexes, and biodistribution studies in mice indicated a fast rate of blood clearance and high rate of total radioactivity excretion, occurring primarily through the renal-urinary pathway. Based on the framework of the asymmetric chelators, the novel bifunctional ligands 3,5-Me(2)-pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(6)) and pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(7)) have been synthesized and their coordination chemistry toward (NEt(4))(2)[ReBr(3)(CO)(3)] (1) has been explored. The resulting complexes, fac-[Re(CO)(3)(k(3)-L)]Br (L(6)(7), L(7)(8)), contain tridentate ancillary ligands that are coordinated to the metal center through the pyrazolyl and amine nitrogen atoms, as observed for the other related building blocks. L(6) and L(7) were coupled to a glycylglycine ethyl ester dipeptide, and the resulting functionalized ligands were used to prepare the model complexes fac-[Re(CO)(3)(kappa(3)-3,5-Me(2)-pz(CH(2))(2)N(glygly)(CH(2))(2)NH(2))](+) (9/9a) and fac-[Re(CO)(3)(kappa(3)-pz(CH(2))(2)N(CH(2))(3)(glygly)(CH(2))(2)NH(2))](+) (10/10a) (M = Re, (99m)Tc). These small conjugates have been fully characterized and are reported herein. On the basis of the in vitro/in vivo behavior of the model complexes (2a-6a, 9a, 10a), we chose to evaluate the in vitro/in vivo biological behavior of a new tumor-seeking Bombesin pyrazolyl conjugate, [(L(6))-G-G-G-Q-W-A-V-G-H-L-M-NH(2)], that has been labeled with the [(99m)Tc(CO)(3)](+) metal fragment. Stability, in vitro cell binding assays, and pharmacokinetics studies in normal mice are reported herein.  相似文献   

4.
5.
The novel anticancer drug ([[trans-PtCl(NH(3))(2)](2)-mu-[trans-Pt(NH(3))(2)(NH(2)(CH(2))(6)NH(2))(2)]](NO(3))(4)) (BBR3464, 1,0,1/t,t,t, TPC) forms a 1,4-interstrand cross-linked adduct with the self-complementary DNA octamer 5'-d(ATG*TACAT)(2)-3', with the two platinum atoms coordinated in the major groove at N7 positions of guanines four base pairs apart on opposite DNA strands [Y. Qu, N.J. Scarsdale, M.-C. Tran, N. Farrell, J. Biol. Inorg. Chem. 8 (2003) 19-28]. The structure of the identical cross-link formed by the dinuclear [[trans-PtCl(NH(3))(2)](2)-mu-NH(2)(CH(2))(6)NH(2)]](NO(3))(2) (BBR3005, 1,1/t,t, DPC) was examined for comparison. The adduct was characterized and analyzed by MS, UV and NMR spectroscopy. NMR analysis of the adduct shows platination of the unique guanine residues. The strong H8/H1' intraresidue cross-peaks observed for all purine residues (A1, G3, A5 and A7) are consistent with a syn-conformation of the nucleoside unit in all cases. Thus, the structure resembles closely that formed by the trinuclear compound. Further confirmation of this similarity comes from the increase in melting temperature (66 degrees for DPC, 60 degrees for TPC, 22 degrees for free oligonucleotide). Since DNA is the principal target in vivo for these Pt cross-linking agents, the unique structural perturbations induced by these cross-links may be related to the increased cytotoxicity and antitumor activity of polynuclear platinum compounds as compared to cisplatin (cis-DDP). The similarity in the structures suggests opportunities to "deliver" the cross-link in a more efficient manner than the current clinically tested drug.  相似文献   

6.
Apparent Ca(2+)-binding constant (K(app)) of Caenorhabditis elegans troponin C (CeTnC) was determined by a fluorescence titration method. The K(app) of the N-domain Ca(2+)-binding site of CeTnC was 7.9+/-1.6 x 10(5) M(-1) and that of the C-domain site was 1.2+/-0.6 x 10(6) M(-1), respectively. Mg(2+)-dependence of the K(app) showed that both Ca(2+)-binding sites did not bind competitively Mg(2+). The Ca(2+) dissociation rate constant (k(off)) of CeTnC was determined by the fluorescence stopped-flow method. The k(off) of the N-domain Ca(2+)-binding site of CeTnC was 703+/-208 s(-1) and that of the C-domain site was 286+/-33 s(-1), respectively. From these values we could calculate the Ca(2+)-binding rate constant (k(on)) as to be 5.6+/-2.8 x 10(8) M(-1) s(-1) for the N-domain site and 3.4+/-2.1 x 10(8) M(-1) s(-1) for the C-domain site, respectively. These results mean that all Ca(2+)-binding sites of CeTnC are low affinity, fast dissociating and Ca(2+)-specific sites. Evolutional function of TnC between vertebrate and invertebrate and biological functions of wild type and mutant CeTnCs are discussed.  相似文献   

7.
Degradation of the reduced pyridine nucleotides NMNH and NADH by HOCl involves two distinct stages: a fast reaction, k = 4.2 x 10(5) M(-1) s(-1), leads to generation of stable pyridine products (Py/Cl) with a strong absorption band at 275 nm (epsilon = 12.4 x 10(3) M(-1) cm(-1) in the case of NMNH); secondarily, a subsequent reaction of HOCl, k = 3.9 x 10(3) M(-1) s(-1), leads to a complete loss of the aromatic absorption band of the pyridine ring. HOBr and HOI(I(2)) react similarly. Apparent rate constants of the primary reactions of HOX species with NMNH at pH 7.2 increase in the order HOCl (3 x 10(5) M(-1) s(-1)) < HOBr( approximately 4 x 10(6) M(-1) s(-1)) < HOI(I(2))( approximately 6.5 x 10(7) M(-1) s(-1)). HOBr reacts fast also with the primary product Py/Br, k approximately 9 x 10(5) M(-1) s(-1), while the reactions of HOI and I(2) with Py/I are slower, approximately 1.4 x 10(3) M(-1) s(-1) and >6 x 10(3) M(-1) s(-1), respectively. Halogenation of the amide group of NMN(+) by HOX species is many orders of magnitude slower than oxidation of NMNH. Taurine inhibits HOCl-induced oxidation of NADH, but HOBr-induced oxidation is not inhibited because the taurine monobromamine rapidly oxidizes NADH, and oxidation by HOI(I(2)) is not inhibited because taurine is inert toward HOI(I(2)). Also sulfur compounds (GSH, GSSG, and methionine) are less efficient in protecting NADH against oxidation by HOBr and HOI(I(2)) than against oxidation by HOCl. The results suggest that reactions of HOBr and HOI(I(2)) in a cellular environment are much more selectively directed toward irreversible oxidation of NADH than reactions of HOCl. It is noteworthy that the rather inert N-chloramines react with iodide to generate HOI(I(2)), i.e., the most reactive and selective oxidant of reduced pyridine nucleotides. NMR investigations show that the primary stable products of the reaction between NMNH and HOCl are various isomeric chlorohydrins originating from a nonstereospecific electrophilic addition of HOCl to the C5&dbond;C6 double bond of the pyridine ring. The primary products (Py/X) of NMNH all exhibit similar absorption bands around 275 nm and are hence likely to result from analogous addition of HOX to the C5&dbond;C6 bond of the pyridine ring. Since the Py/X species are stable and inert toward endogeneous reductants like ascorbate and GSH, they may generally be useful markers for assessing the contribution of hypohalous acids to inflammatory injury.  相似文献   

8.
Xia Z  Azurmendi HF  Mildvan AS 《Biochemistry》2005,44(46):15334-15344
The MutT pyrophosphohydrolase, in the presence of Mg2+, catalyzes the hydrolysis of nucleoside triphosphates by nucleophilic substitution at Pbeta, to yield the nucleotide and PP(i). The best substrate for MutT is the mutagenic 8-oxo-dGTP, on the basis of its Km being 540-fold lower than that of dGTP. Product inhibition studies have led to a proposed uni-bi-iso kinetic mechanism, in which PP(i) dissociates first from the enzyme-product complex (k3), followed by NMP (k4), leaving a product-binding form of the enzyme (F) which converts to the substrate-binding form (E) in a partially rate-limiting step (k5) [Saraswat, V., et al. (2002) Biochemistry 41, 15566-15577]. Single- and multiple-turnover kinetic studies of the hydrolysis of dGTP and 8-oxo-dGTP and global fitting of the data to this mechanism have yielded all of the nine rate constants. Consistent with an "iso" mechanism, single-turnover studies with dGTP and 8-oxo-dGTP hydrolysis showed slow apparent second-order rate constants for substrate binding similar to their kcat/Km values, but well below the diffusion limit (approximately 10(9) M(-1) s(-1)): k(on)app = 7.2 x 10(4) M(-1) s(-1) for dGTP and k(on)app = 2.8 x 10(7) M(-1) s(-1) for 8-oxo-dGTP. These low k(on)app values are fitted by assuming a slow iso step (k5 = 12.1 s(-1)) followed by fast rate constants for substrate binding: k1 = 1.9 x 10(6) M(-1) s(-1) for dGTP and k1 = 0.75 x 10(9) M(-1) s(-1) for 8-oxo-dGTP (the latter near the diffusion limit). With dGTP as the substrate, replacing Mg2+ with Mn2+ does not change k1, consistent with the formation of a second-sphere MutT-M2+-(H2O)-dGTP complex, but slows the iso step (k5) 5.8-fold, and its reverse (k(-5)) 25-fold, suggesting that the iso step involves a change in metal coordination, likely the dissociation of Glu-53 from the enzyme-bound metal so that it can function as the general base. Multiple-turnover studies with dGTP and 8-oxo-dGTP show bursts of product formation, indicating partially rate-limiting steps following the chemical step (k2). With dGTP, the slow steps are the chemical step (k2 = 10.7 s(-1)) and the iso step (k5 = 12.1 s(-1)). With 8-oxo-dGTP, the slow steps are the release of the 8-oxo-dGMP product (k4 = 3.9 s(-1)) and the iso step (k5 = 12.1 s(-1)), while the chemical step is fast (k2 = 32.3 s(-1)). The transient kinetic studies are generally consistent with the steady state kcat and Km values. Comparison of rate constants and free energy diagrams indicate that 8-oxo-dGTP, at low concentrations, is a better substrate than dGTP because it binds to MutT 395-fold faster, dissociates 46-fold slower, and has a 3.0-fold faster chemical step. The true dissociation constants (KD) of the substrates from the E-form of MutT, which can now be obtained from k(-1)/k1, are 3.5 nM for 8-oxo-dGTP and 62 microM for dGTP, indicating that 8-oxo-dGTP binds 1.8 x 10(4)-fold tighter than dGTP, corresponding to a 5.8 kcal/mol lower free energy of binding.  相似文献   

9.
The kinetic properties of placental glucose-6-phosphate dehydrogenase were studied, since this enzyme is expected to be an important component of the placental protection system. In this capacity it is also very important for the health of the fetus. The placental enzyme obeyed "Rapid Equilibrium Ordered Bi Bi" sequential kinetics with K(m) values of 40+/-8 microM for glucose-6-phosphate and 20+/-10 microM for NADP. Glucose-6-phosphate, 2-deoxyglucose-6-phosphate and galactose-6-phosphate were used with catalytic efficiencies (k(cat)/K(m)) of 7.4 x 10(6), 4.89 x 10(4) and 1.57 x 10(4) M(-1).s(-1), respectively. The K(m)app values for galactose-6-phosphate and for 2-deoxyglucose-6-phosphate were 10+/-2 and 0.87+/-0.06 mM. With galactose-6-phosphate as substrate, the same K(m) value for NADP as glucose-6-phosphate was obtained and it was independent of galactose-6-phosphate concentration. On the other hand, when 2-deoxyglucose-6-phosphate used as substrate, the K(m) for NADP decreased from 30+/-6 to 10+/-2 microM as the substrate concentration was increased from 0.3 to 1.5 mM. Deamino-NADP, but not NAD, was a coenzyme for placental glucose-6-phosphate dehydrogenase. The catalytic efficiencies of NADP and deamino-NADP (glucose-6-phosphate as substrate) were 1.48 x 10(7) and 4.80 x 10(6) M(-1)s(-1), respectively. With both coenzymes, a hyperbolic saturation and an inhibition above 300 microM coenzyme concentration, was observed. Human placental glucose-6-phosphate dehydrogenase was inhibited competitively by 2,3-diphosphoglycerate (K(i)=15+/-3 mM) and NADPH (K(i)=17.1+/-3.2 microM). The small dissociation constant for the G6PD:NADPH complex pointed to tight enzyme:NADPH binding and the important role of NADPH in the regulation of the pentose phosphate pathway.  相似文献   

10.
Complex [Cu(dpq)(2)(H(2)O)](ClO(4))(2).H(2)O (1), where dpq is dipyrido-[3,2-D:2',3'-f]-quinoxaline, has been prepared by reacting copper(II) perchlorate hexahydrate with dpq in methanol and structurally characterized. The complex crystallizes in the triclinic space group P-1 with the unit cell parameters a=8.646(2) A, b=12.290(5) A, c=14.283(4) A, alpha=94.01(2) degrees, beta=91.69(2) degrees,gamma=101.60 (3) degrees, V=1481.7(8) A(3) and Z=2. The structure, refined to R=0.0505 and R(w)=0.1441 for 5212 reflections with I>2sigma (I) using 440 parameters, shows the presence of a CuN(4)O chromophore in an axially compressed distorted trigonal-bipyramidal structure. The Cu-N distances lie in the range 1.969(3)-2.103(3) A. The Cu-OH(2) distance is 2.145(3) A. The complex is one-electron paramagnetic and exhibits a visible spectral d-d band at 718 nm in MeCN. It shows a quasi-reversible cyclic voltammetric response at 0.091 V (DeltaE(p)=229 mV) at 50 mV s(-1) in MeCN-0.1 M TBAP for the Cu(II)/Cu(I) couple. In 50 mM Tris-HCl/0.1 M KCl buffer-DMF mixture (1:4 v/v, pH 7.2), the couple appears at 0.089 V versus SCE. The complex undergoes facile reduction with sodium ascorbate in an aqueous DMF mixture (4:1 v/v) to form an unstable brown Cu(I) species (lambda(max)=440 nm, epsilon=7480 M(-1) cm(-1)) which converts to 1 on exposure to air giving a turnover frequency of ca. 400. Binding studies revealed that 1 is an efficient binder to calf thymus DNA. Complex 1 on reaction with supercoiled (SC) DNA in presence of ascorbic acid in a 50 mM Tris-HCl/50 mM NaCl buffer (pH 7.2) shows nuclease activity which is 4.5 times greater than that of the phen analogue.  相似文献   

11.
In addition to its catalytic domain, phosphoinsositide-dependent protein kinase-1 (PDK1) contains a C-terminal pleckstrin homology (PH) domain, which binds the membrane-bound phosphatidylinositol (3,4,5)-triphosphate [PI(3,4,5)P3] second messenger. Here, we report in vitro kinetic, phosphopeptide mapping, and oligomerization studies that address the role of the PH domain in regulating specific autophosphorylation events, which are required for PDK1 catalytic activation. First, 'inactive' unphosphorylated forms of N-terminal His6 tagged full length (His6-PDK1) and catalytic domain constructs [His6-PDK1(Delta PH)] were generated by treatment with Lambda protein phosphatase (lambda PP). Reconstitution of lambda PP-treated His6-PDK1(Delta PH) catalytic activity required activation loop Ser-241 phosphorylation, which occurred only upon trans-addition of 'active' PDK1 with an apparent bimolecular rate constant of (app)k1(S241) = 374+/-29 M(-1) s(-1). In contrast, full length lambda PP-treated His6-PDK1 catalyzed Ser-241 cis-autophosphorylation with an apparent first-order rate constant of (app)k1(S241) = (5.0+/-1.5) x 10(-4) s(-1) but remained 'inactive'. Reconstitution of lambda PP-treated His(6)-PDK1 catalytic activity occurred only when autophosphorylated in the presence of PI(3,4,5)P3 containing vesicles. PI(3,4,5)P3 binding to the PH domain activated apparent first-order Ser-241 autophosphorylation by 20-fold [(app)k1(S241) = (1.1+/-0.1) x 10(-2) s(-1)] and also promoted biphasic Thr-513 trans-autophosphorylation [(app)k2(T513) = (4.9+/-1.1) x 10(2) M(-1) s(-1) and(app)k3(T513) = (1.5+/-0.2) x 10(3) M(-1) s(-1)]. The results of mutagenesis studies suggest that Thr-513 phosphorylation may cause dissociation of autoinhibitory contacts formed between the contiguous regulatory PH and catalytic kinase domains.  相似文献   

12.
Double-stranded (ds) oligodeoxynucleotides (29mers) containing an O6-ethylguanine (O6-EtGua) flanked 5' and 3' by different bases (5'..TGT..3'; 5'..CGG..3', 5'..GGT..3'; 5'..GGG..3'; 5'..GGA..3') were synthesized to investigate the binding and repair characteristics of recombinant human O6-alkylguanine-DNA alkyltransferase (AT) in vitro. The apparent association constant (KA(app)) of AT to the oligomers and the repair rate constant for O6-EtGua (k) respectively, were determined by gel retardation and a monoclonal antibody-based filter binding assay. When ds- or single-stranded (ss) oligomers with or without O6-EtGua were used, no major differences in KA(app) values were observed with either substrate: KA(app) values for native AT were 7.1 and 8.4 x 10(5) M(-1) respectively, for unmodified and [O6-EtGua]-containing ds-oligomers. The corresponding values for ss-oligomers were 1.0 and 4.9 x 10(5) M(-1). The N-terminal first 56 amino acids of AT only exert a limited influence on DNA binding; the KA(app) values for an N-terminally truncated AT protein (1.1 x 10(5) M(-1)) and native AT were of the same order. Moreover, KA(app) was hardly affected by Cys(145)-methylated AT (2.0 x 10(5) M(-1)). The k-values (6.5-11.5 x 10(6) M(-1)s(-1)) were not significantly dependent on nucleotide sequence. k-values of 5.3 and 4.0 x 10(6) M(-1)s(-1) respectively, were obtained with the N-terminally truncated AT protein and for repair of the postreplicative mispair [O6-EtGua]: T by native AT. The low KA(app), the negligible influence on O6 of ethylation, and the minor modulation KA(app) and k by varying the bases flanking O6-EtGua, all indicate that the binding of AT to DNA is non-specific and mediated mainly by ionic interactions [reduced KA(app) and k-values at increased ionic strength]. Surplus DNA reduces the rate of O6-EtGua repair in ds-oligomers by competitive binding of AT molecules. The reaction mechanism of AT with DNA in vivo requires further investigation.  相似文献   

13.
Ion permeation and conduction were studied using whole-cell recordings of the M-current (I(M)) and delayed rectifier (IDR), two K+ currents that differ greatly in kinetics and modulation. Currents were recorded from isolated bullfrog sympathetic neurons with 88 mM [K+]i and various external cations. Selectivity for extracellular monovalent cations was assessed from permeability ratios calculated from reversal potentials and from chord conductances for inward current. PRb/PK was near 1.0 for both channels, and GRb/GK was 0.87 +/- 0.01 for IDR but only 0.35 +/- 0.01 for I(M) (15 mM [Rb+]o or [K+]o). The permeability sequences were generally similar for I(M) and IDR: K+ approximately Rb+ > NH4+ > Cs+, with no measurable permeability to Li+ or CH3NH3+. However, Na+ carried detectable inward current for IDR but not I(M). Nao+ also blocked inward K+ current for IDR (but not IM), at an apparent electrical distance (delta) approximately 0.4, with extrapolated dissociation constant (KD) approximately 1 M at 0 mV. Much of the instantaneous rectification of IDR in physiologic ionic conditions resulted from block by Nao+. Extracellular Cs+ carried detectable inward current for both channel types, and blocked I(M) with higher affinity (KD = 97 mM at 0 mV for I(M), KD) approximately 0.2 M at 0 mV for IDR), with delta approximately 0.9 for both. IDR showed several characteristics reflecting a multi-ion pore, including a small anomalous mole fraction effect for PRb/PK, concentration-dependent GRb/GK, and concentration- dependent apparent KD's and delta's for block by Nao+ and Cso+. I(M) showed no clear evidence of multi-ion pore behavior. For I(M), a two- barrier one-site model could describe permeation of K+ and Rb+ and block by Cso+, whereas for IDR even a three-barrier, two-site model was not fully adequate.  相似文献   

14.
A novel amperometric biosensor for hydrogen peroxide (H(2)O(2)) was developed by entrapping horseradish peroxidase (HRP) in a new ormosil composite doped with ferrocene monocarboxylic acid-bovine serum albumin conjugate and multiwall carbon nanotubes (MWNTs). The ormosil was prepared using 3-(aminopropyl)triethoxysilane and 2-(3,4 epoxycyclohexyl)-ethyltrimethoxy silane as monomers. The encapsulated conjugate showed excellent electrochemistry and acted as an electron transfer mediator. The presence of MWNTs improved the conductivity of the composite film. This matrix showed a biocompatible microenvironment for retaining the native activity of the entrapped HRP and a very low mass transport barrier to the substrate, which provided a fast amperometric response to H(2)O(2). The proposed H(2)O(2) biosensor exhibited a linear range of 0.02-4.0 mM with a detection limit of 5.0 microM (S/N = 3) and a K(M)(app) value of 2.0 mM. It could be used for flow injection analysis of hydrogen peroxide with a liner range from 0.02 to 4.5 mM, sensitivity of 0.042 microA/mM and analytical time of 20 s per sample. This biosensor possessed good analytical performance and storage stability.  相似文献   

15.
The hammerhead cleavage reaction in monovalent cations   总被引:10,自引:3,他引:7       下载免费PDF全文
Recently, Murray et al. (Chem Biol, 1998, 5:587-595) found that the hammerhead ribozyme does not require divalent metal ions for activity if incubated in high (> or =1 M) concentrations of monovalent ions. We further characterized the hammerhead cleavage reaction in the absence of divalent metal. The hammerhead is active in a wide range of monovalent ions, and the rate enhancement in 4 M Li+ is only 20-fold less than that in 10 mM Mg2+. Among the Group I monovalent metals, rate correlates in a log-linear manner with ionic radius. The pH dependence of the reaction is similar in 10 mM Mg2+, 4 M Li+, and 4 M Na+. The exchange-inert metal complex Co(NH3)3+ also supports substantial hammerhead activity. These results suggest that a metal ion does not act as a base in the reaction, and that the effects of different metal ions on hammerhead cleavage rates primarily reflect structural contributions to catalysis.  相似文献   

16.
Eliot AC  Kirsch JF 《Biochemistry》2002,41(11):3836-3842
The active sites of the homologous pyridoxal phosphate- (PLP-) dependent enzymes 1-aminocyclopropane-1-carboxylate (ACC) synthase and aspartate aminotransferase (AATase) are almost entirely conserved, yet the pK(a)'s of the two internal aldimines are 9.3 and 7.0, respectively, to complement the substrate pK(a)'s (S-adenosylmethionine pK(a) = 7.8 and aspartate pK(a) = 9.9). This complementation is required for maximum enzymatic activity in the physiological pH range. The most prominent structural difference in the active site is that Ile232 of ACC synthase is replaced by alanine in AATase. The I232A mutation was introduced into ACC synthase with a resulting 1.1 unit decrease (from 9.3 to 8.2) in the aldimine pK(a), thus identifying Ile232 as a major determinant of the high pK(a) of ACC synthase. The mutation also resulted in reduced k(cat) (0.5 vs 11 s(-1)) and k(cat)/K(m) values (5.0 x 10(4) vs 1.2 x 10(6) M(-1) s(-1)). The effect of the mutation is interpreted as the result of shortening of the Tyr233-PLP hydrogen bond. Addition of the Y233F mutation to the I232A ACC synthase to generate the double mutant I232A/Y233F raised the pK(a) from 8.2 to 8.8, because the Y233F mutation eliminates the hydrogen bond between that residue and PLP. The introduction of the retro mutation A224I into AATase raised the aldimine pK(a) of that enzyme from 6.96 to 7.16 and resulted in a decrease in single-turnover k(max) (108 vs 900 s(-1) for aspartate) and k(max)/K(m)(app) (7.5 x 10(4) vs 3.8 x 10(5) M(-1) s(-1)) values. The distance from the pyridine nitrogen of the cofactor to a conserved aspartate residue is 2.6 A in AATase and 3.8 A in ACC synthase. The D230E mutation introduced into ACC synthase to close this distance increases the aldimine pK(a) from 9.3 to 10.0, as would be predicted from a shortened hydrogen bond.  相似文献   

17.
Recombinant full-length human procathepsin F, produced in the baculovirus expression system, was partially processed during the purification procedure to a form lacking the N-terminal cystatin-like domain and activated with pepsin. Active cathepsin F efficiently hydrolyzed Z-FR-MCA (kcat/Km=106 mM(-1) s(-1)) and Bz-FVR-MCA (kcat/Km=8 mM(-1) s(-1)), whereas hydrolysis of Z-RR-MCA was very slow (kcat/Km<0.2 mM(-1) s(-1)). Cathepsin F was rapidly and tightly inhibited by cystatin C, chicken cystatin and equistatin with Ki values in the subnanomolar range (0.03-0.47 nM), whereas L-kininogen was a less strong inhibitor of the enzyme (Ki=4.7 nM). Stefin A inhibited cathepsin F slowly (kass=1.6 x 10(5) M(-1) s(-1)) and with a lower affinity (Ki=25 nM). These data suggest that cathepsin F differs from other related endopeptidases by considerably weaker inhibition by stefins.  相似文献   

18.
The Cu(I)-phenolate complexes (1)LCu and (2)LCu and the Cu(I)-phenol complex [H(2)LCu(CNC(6)H(3)Me(2))]BArF(4) were prepared and structurally characterized by X-ray crystallography, where (1)L(-) and (2)L(-) are ligands comprised of a 2,4-di- tert-butylphenolate linked to 1-isopropyl-1,5-diazacyclooctane or 1,4-diisopropyl-1,4,7-triazacyclononane, respectively. The reduced galactose oxidase (GAO) structural models (1)LCu and (2)LCu were found to be highly reactive with O(2), and through combined stopped-flow kinetic and EPR, UV-vis, and resonance Raman spectroscopic studies of the oxygenation of (2)LCu at low temperature, new intermediates relevant to those postulated for the active site oxidation step of the GAO catalytic cycle were identified. The oxygenation was shown by kinetics experiments to proceed via initial binding of O(2) to yield a green, unusually thermodynamically stable 1:1 adduct, (2)LCu(O(2)). Symmetric (eta(2)) binding of a superoxo ligand was indicated by oxygen-isotope-sensitive features in resonance Raman spectra obtained in batch experiments; peaks at nu((16)O(2))=1120 cm(-1), nu((18)O(16)O)=1093 cm(-1), and nu((18)O(2))=1058 cm(-1) were assigned as O-O stretching vibrations. These data represent the first experimental evidence for such superoxide coordination in complexes of tetradentate tripodal ligands and provide new precedent for how O(2) may bind at the reduced GAO active site. The 1:1 Cu/O(2) adduct subsequently evolves into a metastable purple species that is only observable under conditions of substoichiometric O(2). The kinetics of formation of this transient species are second order overall (rate= k'(2)[(2)LCu(O(2))][(2)LCu]). It exhibits an absorption band with lambda(max)=565 nm (epsilon=17900 M(-1) cm(-1)) and multiple oxygen-isotope-sensitive nu(Cu-O) and nu(O-O) features in the respective regions 500-550 cm(-1) and 700-850 cm(-1) in Raman spectra, with excitation-wavelength-dependent intensities that correlate with the 565 nm absorption feature. On the basis of the combined data available, the presence of multiple isomeric peroxodicopper species in the transient purple solution is postulated.  相似文献   

19.
The acid-base properties of γ-L-glutamyl-L-cysteinyl-glycine (glutathione, GSH) were determined by potentiometry (ISE-H(+), glass electrode) in pure NaI((aq)) and in NaCl((aq))/MgCl(2(aq)), and NaCl((aq))/CaCl(2(aq)) mixtures, at T = 298.15 K and different ionic strengths (up to I(c) ~ 5.0 mol L(-1)). In addition, the activity coefficients of glutathione were also determined by the distribution method at the same temperature in various ionic media (LiCl((aq)), NaCl((aq)), KCl((aq)), CsCl((aq)), MgCl(2(aq)), CaCl(2(aq)), NaI((aq))). The results obtained were also used to calculate the Specific ion Interaction Theory (SIT) and Pitzer coefficients for the dependence on medium and ionic strength of glutathione species, as well as the formation constants of weak Mg(j)H( i )(GSH)((i+2j-3)) and Ca(j)H(i)(GSH)((i+2j-3)) complexes. Direct calorimetric titrations were also carried out in pure NaCl((aq)) and in NaCl((aq))/CaCl(2(aq)) mixtures at different ionic strengths (0.25 ≤ I (c )/mol L(-1) ≤ 5.0) in order to determine the enthalpy changes for the protonation and complex formation equilibria in these media at T = 298.15 K. Results obtained are useful for the definition of glutathione speciation in any aqueous media containing the main cations of natural waters and biological fluids, such as Na(+), K(+), Mg(2+), and Ca(2+). Finally, this kind of systematic studies, where a series of ionic media (e.g., all alkali metal chlorides) is taken into account in the determination of various thermodynamic parameters, is useful for the definition of some trends in the thermodynamic behavior of glutathione in aqueous solution.  相似文献   

20.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  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