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 828 毫秒
1.
A strategic assessment of the contributions of two active-site hydrogen bonds in the binding of nicotinate to recombinant ferric soybean leghaemoglobin a (rLb) was carried out by mutagenic replacement of the hydrogen-bonding residues (H61A and Y30A variants) and by complementary chemical substitution of the carboxylate functionality on the nicotinate ligand. Dissociation constants, Kd (pH 5.5, mu = 0.10 M, 25.0 +/- 0.1 degrees C), for binding of nicotinate to ferric rLb, H61A and Y30A were 1.4 +/- 0.3 microM, 19 +/- 1 microM and 11 +/- 1 microM, respectively; dissociation constants for binding of nicotinamide were, respectively, 38 +/- 1 mM, 50 +/- 2 mM and 12 +/- 1 mM, and for binding of pyridine were 260 +/- 50 microM, 4.5 +/- 0.5 microM and 66 +/- 8 microM, respectively. Binding of cyanide and azide to the H61A and Y30A variants was unaffected by the mutations. The pH-dependence of nicotinate binding for rLb and Y30A was consistent with a single titration process (pKa values 6.9 +/- 0.1 and 6.7 +/- 0.2, respectively); binding of nicotinate to H61A was independent of pH. Reduction potentials for the rLb and rLb-nicotinate derivatives were 29 +/- 2 mV (pH 5.40, 25.0 degrees C, mu = 0.10 M) and - 65 +/- 2 mV (pH 5.42, 25.0 degrees C, mu = 0.10 M), respectively. The experiments provide a quantitative assessment of the role of individual hydrogen bonds in the binding process, together with a definitive determination of the pKa of His61 and unambiguous evidence that titration of His61 controls binding in the neutral to alkaline region.  相似文献   

2.
We have exploited the intrinsic conformational flexibility of leghemoglobin to reengineer the heme active site architecture of the molecule by replacement of the mobile His61 residue with tyrosine (H61Y variant). The electronic absorption spectrum of the ferric derivative of H61Y is similar to that observed for the phenolate derivative of the recombinant wild-type protein (rLb), consistent with coordination of Tyr61 to (high-spin) iron. EXAFS data clearly indicate a 6-coordinate heme geometry and a Fe-O bond length of 185pm. MCD and EPR spectroscopies are consistent with this assignment and support ligation by an anionic (tyrosinate) group. The alteration in heme ligation leads to a 148mV decrease in the reduction potential for H61Y (-127+/-5mV) compared to rLb and destabilisation of the functional oxy-derivative. The results are discussed in terms of our wider understanding of other heme proteins with His-Tyr ligation.  相似文献   

3.
Catalytic oxidation of p-cresol by ascorbate peroxidase   总被引:2,自引:0,他引:2  
Transient and steady state kinetics, together with a range of chromatographic and spectroscopic techniques, have been used to establish the mechanism and the products of the H(2)O(2)-dependent oxidation of p-cresol by ascorbate peroxidase (APX). HPLC, GC-MS, and NMR analyses are consistent with the formation of 2, 2'-dihydroxy-5,5'-dimethylbiphenyl (II) and 4alpha,9beta-dihydro-8, 9beta-dimethyl-3(4H)-dibenzofuranone (Pummerer's ketone, III) as the major products of the reaction. In the presence of cumene hydroperoxide, two additional products were observed which, from GC and MS analyses, were shown to be 1,1-dimethylbenzylalcohol (IV) and bis-(1-methyl-1-phenyl-ethyl)-peroxide (V). The product ratio II:III was dependent on enzyme concentration: at low concentrations Pummerer's ketone (III) predominates and at high concentrations formation of the biphenyl compound (II) is favored. Steady-state data showed a sigmoidal dependence on [p-cresol] that was consistent with the presence of 2.01 +/- 0.15 binding sites for the substrate (25.0 degrees C, sodium phosphate, pH 7.0, mu = 2.2 mM) and independent of ionic strength in the range 2.2-500 mM. Single turnover kinetic experiments (pH 7.0, 5.0 degrees C, mu = 0.10 M) yielded second-order rate constants for Compound I reduction by p-cresol, k(2), of 5.42 +/- 0.10 x 10(5) M(-1) s(-1), respectively. Rate-limiting reduction of Compound II by p-cresol, k(3), showed saturation kinetics, giving values for K(d) = 1.54 +/- 0.12 x 10(-3) M and k(3) = 18.5 +/- 0.7 s(-1). The results are discussed in the more general context of APX-catalyzed aromatic oxidations.  相似文献   

4.
The thermodynamic and catalytic properties of flavocytochrome c3 from Shewanella frigidimarina have been studied using a combination of protein film voltammetry and solution methods. As measured by solution kinetics, maximum catalytic efficiencies for fumarate reduction (kcat/Km = 2.1 x 10(7) M-1 s-1 at pH 7.2) and succinate oxidation (kcat/Km = 933 M-1 s-1 at pH 8.5) confirm that flavocytochrome c3 is a unidirectional fumarate reductase. Very similar catalytic properties are observed for the enzyme adsorbed to monolayer coverage at a pyrolytic graphite "edge" electrode, thus confirming the validity of the electrochemical method for providing complementary information. In the absence of fumarate, the adsorbed enzyme displays a complex envelope of reversible redox signals which can be deconvoluted to yield the contributions from each active site. Importantly, the envelope is dominated by the two-electron signal due to FAD [E degrees ' = -152 mV vs the standard hydrogen electrode (SHE) at pH 7.0 and 24 degrees C] which enables quantitative examination of this center, the visible spectrum of which is otherwise masked by the intense absorption bands due to the hemes. The FAD behaves as a cooperative two-electron center with a pH-dependent reduction potential that is modulated (pKox at 6.5) by ionization of a nearby residue. In conjunction with the kinetic pKa values determined for the forward and reverse reactions (7.4 and 8.6, respectively), a mechanism for fumarate reduction, incorporating His365 and an anionic form of reduced FAD, is proposed. The reduction potentials of the four heme groups, estimated by analysis of the underlying envelope, are -102, -146, -196, and -238 mV versus the SHE at pH 7.0 and 24 degrees C and are comparable to those determined by redox potentiometry.  相似文献   

5.
A nonsporulating fungus isolated from dioxine-containing tropical soils forms cellobiose dehydrogenase, when grown in media supplemented by a source of cellulose. The enzyme purified to homogeneity by SDS-PAGE (yield, 43%) had an M(r) of 95 kDa; its pH optimum was in the range 5.5-7.0; more than 50% activity was retained at pH 4.0-8.0 (citrate-phosphate buffer). The absorption spectrum of the enzyme in the visible range had the characteristic appearance of flavocytochrome proteins. Cellobiose dehydrogenase oxidized cellobiose and lactose (the respective K(M) values at pH 6.0 equaled 4.5 +/- 1.5 and 56 microM) in the presence of dichlorophenolindophenol (K(M) app = 15 +/- 3 microM at pH 6.0) taken as an electron acceptor. Other sugars were barely if at all oxidized by the enzyme. Neither ethyl-beta-D-cellobioside, heptobiose, nor chitotriose inhibited the enzymatic oxidation of lactose, even under the conditions of 100-fold molar excess. The enzyme was weakly inhibited by sodium azide dichlorophenolindophenol reduction and exhibited affinity to amorphous cellulose. At 55 degrees C and pH 6.0 (optimum stability), time to half-maximum inactivation equaled 99 min. The enzyme reduced by cellobiose was more stable than the nonreduced form. Conversely, the presence of an oxidizer (dichlorophenolindophenol) decreased the stability eight times at pH 6.0. In addition, the enzyme acted as a potent reducer of the single-electron acceptor cytochrome c3+ (K(M) app = 15 microM at pH 6.0).  相似文献   

6.
To further the understanding of the biological importance of metal-binding by avian prion proteins, we have investigated the affinity and selectivity of peptides Hx1 [Ac-HNPGYP-nh] and Hx2 [Ac-NPGYPHNPGYPH-nh] with a range of physiological metals via electrospray ionization mass spectrometry and tyrosine fluorescence emission spectroscopy. Both the hexamer Hx1 and the "dimer" peptide Hx2 bind only one equivalent of Cu(II), although only the latter peptide binds copper with significant affinity (Hx1 K(d)=150+/-35 microM; Hx2 K(d)=1.07+/-0.78 microM, pH 7.0 in 3-(N-morpholino)propanesulfonic acid (MOPS) buffer). Both peptides are selective for Cu(II) over divalent Ca, Co, Mg, Mn, Ni, and Zn. Cyclic voltammetry was used to estimate Cu(II/I) solution potentials at pH 6.8, which were very similar for the two peptides (CuHx1 E degrees'=+350 mV, CuHx2 E degrees'=+320 mV vs. normal hydrogen electrode). These results suggest similar binding modes for the two peptides, and relative stabilization of Cu(I) relative to similar His-Gly-rich peptides in the literature.  相似文献   

7.
The Mg2+-ATPase activity of myosin and its subfragment 1 (ATP phosphohydrolase, EC 3.6.1.3) always followed normal Michaelis-Menten kinetics for ATP concentrations less than 10 microM. The average Km values at pH 7.4 and 25 degrees C are 0.33 +/- 0.04 microM for myosin and 0.43 +/- 0.11 microM for subfragment 1. At low salt concentration myosin yields a second hyperbolic increase in Mg2+-ATPase activity as the ATP rises from 10.2 microM to 153 microM: V doubles with a Km of 11 +/- 5 microM. This second low-salt-dependent increase in Mg2+-ATPase activity occurred between pH 6.8 and pH 8.7. It was not affected by the presence of 0.10 M EGTA to remove Ca2+ contamination. Solubilization of the catalytic sites by assaying myosin for ATPase activity in the presence of 0.60 M NaCl or by conversion of myosin to subfragment 1 eliminated the secondary hyperbolic increase. Subfragment 1 has a significantly different pH-activity curve from that of myosin. Subfragment 1 has an activity peak at pH 6.0, a rising activity as the pH goes from 8.7 to 9.8, and a deep activity valley between pH 6.8 and pH 8.4. Myosin has a very shallow trough of activity at pH 6.8 to 8.4, and in 1.0 mM ATP its activity drops as the pH decreases from 6.8 to 6.0. NaCl is a noncompetitive inhibitor of the Mg2+-ATPase activity of myosin and subfragment 1. Myosin has a greater affinity for NaCl (Ki = 0.101 +/- 0.004 M) than does subfragment 1 (Ki = 0.194 +/- 0.009 M).  相似文献   

8.
An intramolecular electron-transfer process has previously been shown to take place between the Cys3--Cys26 radical-ion (RSSR-) produced pulse radiolytically and the Cu(II) ion in the blue single-copper protein, azurin [Farver, O. & Pecht, I. (1989) Proc. Natl Acad. Sci. USA 86, 6868-6972]. To further investigate the nature of this long-range electron transfer (LRET) proceeding within the protein matrix, we have now investigated it in two azurins where amino acids have been substituted by single-site mutation of the wild-type Pseudomonas aeruginosa azurin. In one mutated protein, a methionine residue (Met44) that is proximal to the copper coordination sphere has been replaced by a positively charged lysyl residue ([M44K]azurin), while in the second mutant, another residue neighbouring the Cu-coordination site (His35) has been replaced by a glutamine ([H35Q]azurin). Though both these substitutions are not in the microenvironment separating the electron donor and acceptor, they were expected to affect the LRET rate because of their effect on the redox potential of the copper site and thus on the driving force of the reaction, as well as on the reorganization energies of the copper site. The rate of intramolecular electron transfer from RSSR- to Cu(II) in the wild-type P. aeruginosa azurin (delta G degrees = -68.9 kJ/mol) has previously been determined to be 44 +/- 7 s-1 at 298 K, pH 7.0. The [M44K]azurin mutant (delta G degrees = -75.3 kJ/mol) was now found to react considerably faster (k = 134 +/- 12 s-1 at 298 K, pH 7.0) while the [H35Q]azurin mutant (delta G degrees = -65.4 kJ/mol) exhibits, within experimental error, the same specific rate (k = 52 +/- 11 s-1, 298 K, pH 7.0) as that of the wild-type azurin. From the temperature dependence of these LRET rates the following activation parameters were calculated: delta H++ = 37.9 +/- 1.3 kJ/mol and 47.2 +/- 0.7 kJ/mol and delta S++ = -86.5 +/- 5.8 J/mol.K and -46.4 +/- 4.4 J/mol.K for [H35Q]azurin and [M44K]azurin, respectively. Using the Marcus relation for intramolecular electron transfer and the above parameters we have determined the reorganization energy, lambda and electronic coupling factor, beta. The calculated values fit very well with a through-bond LRET mechanism.  相似文献   

9.
Flavodoxin from Desulfovibrio vulgaris (Hildenborough) has been expressed at a high level (3-4% soluble protein) in Escherichia coli by subcloning a minimal insert carrying the gene behind the tac promoter of plasmid pDK6. The recombinant protein was readily isolated and its properties were shown to be identical to those of the wild-type protein obtained directly from D. vulgaris, with the exception that the recombinant protein lacks the N-terminal methionine residue. Detailed measurements of the redox potentials of this flavodoxin are reported for the first time. The redox potential, E2, for the couple oxidized flavodoxin/flavodoxin semiquinone at pH 7.0 is -143 mV (25 degrees C), while the value for the flavodoxin semiquinone/flavodoxin hydroquinone couple (E1) at the same pH is -440 mV. The effects of pH on the observed potentials were examined; E2 varies linearly with pH (slope = -59 mV), while E1 is independent of pH at high pH values, but below pH 7.5 the potential becomes less negative with decreasing pH, indicating a redox-linked protonation of the flavodoxin hydroquinone. D. vulgaris apoflavodoxin binds FMN very tightly, with a value of 0.24 nM for the dissociation constant (Kd) at pH 7.0 and 25 degrees C, similar to that observed with other flavodoxins. In addition, the apoflavodoxin readily binds riboflavin (Kd = 0.72 microM; 50 mM sodium phosphate, pH 7.0, 5 mM EDTA at 25 degrees C) and the complex is spectroscopically very similar to that formed with FMN. The redox potentials for the riboflavin complex were determined at pH 6.5 (E1 = -262 mV, E2 = -193 mV; 25 degrees C) and are discussed in the light of earlier proposals that charge/charge interactions between different parts of the flavin hydroquinone play a crucial role in determining E1 in flavodoxin.  相似文献   

10.
Vasoconstriction via alpha(2)-receptors is known to be sensitive to acidic pH, but little is known about the pH sensitivity of P2X receptors. ATP is a cotransmitter released with norepinephrine from the sympathetic nerves and causes vasoconstriction via P2X purinergic receptors on vascular smooth muscle. We hypothesized that reductions in pH would attenuate P2X-mediated vasoconstriction in iliofemoral artery rings. Twenty-five rats were killed, and the iliac and femoral arteries were dissected out and placed in modified Krebs-Henseleit buffer. The arteries were cut into 2-mm sections and mounted in an organ tissue bath. Tension (g) was measured during a potassium chloride and norepinephrine challenge (maximal tension). The arteries were then exposed to alpha,beta-methylene ATP (10(-7)-10(-3) M; n = 13) or phenylephrine (10(-7)-10(-4) M; n = 6) with a tissue bath pH of 7.8, 7.4, and 7.0. Dose-response curves were fit with nonlinear regression analysis to calculate the EC(50) and slope. The peak tension with alpha,beta-methylene ATP was lower during pH 7.0 (1.37 +/- 0.09 g) compared with pH 7.8 (1.90 +/- 0.12 g). EC(50) was highest with pH 7.4 (-5.38 +/- 0.18 log M alpha,beta-methylene ATP) and lowest with pH 7.0 (-4.9 +/- 0.10 log M alpha,beta-methylene ATP). The slopes of the dose-response curves were not different. Pyridoxal phosphate-6-azo(benzene-2,4-disulfonic acid) abolished contraction caused by the addition of alpha,beta-methylene ATP (n = 6). There was no effect of pH on phenylephrine dose-response curves. These data indicate that the vasoconstrictor response to alpha,beta-methylene ATP is sensitive to pH and that lower pH attenuates the response of P2X purinergic receptors.  相似文献   

11.
The effect of temperature, pH, free [Mg(2+)], and ionic strength on the apparent equilibrium constant of arginine kinase (EC 2.7.3.3) was determined. At equilibrium, the apparent K' was defined as [see text] where each reactant represents the sum of all the ionic and metal complex species. The K' at pH 7.0, 1.0 mM free [Mg(2+)], and 0. 25 M ionic strength was 29.91 +/- 0.59, 33.44 +/- 0.46, 35.44 +/- 0. 71, 39.64 +/- 0.74, and 45.19 +/- 0.65 (n = 8) at 40, 33, 25, 15, and 5 degrees C, respectively. The standard apparent enthalpy (DeltaH degrees') is -8.19 kJ mol(-1), and the corresponding standard apparent entropy of the reaction (DeltaS degrees') is + 2. 2 J K(-1)mol(-1) in the direction of ATP formation at pH 7.0, free [Mg(2+)] =1.0 mM, ionic strength (I) =0.25 M at 25 degrees C. We further show that the magnitude of transformed Gibbs energy (DeltaG degrees ') of -8.89 kJ mol(-1) is mostly comprised of the enthalpy of the reaction, with 7.4% coming from the entropy TDeltaS degrees' term (+0.66 kJ mol(-1)). Our results are discussed in relation to the thermodynamic properties of its evolutionary successor, creatine kinase.  相似文献   

12.
Recently, we reported that 11,12-epoxyeicosatrienoic acid (11,12-EET) potently activates rat mesenteric arterial ATP-sensitive K(+) (K(ATP)) channels and produces significant vasodilation through protein kinase A-dependent mechanisms. In this study, we tried to further delineate the signaling steps involved in the activation of vascular K(ATP) channels by EETs. Whole cell patch-clamp recordings [0.1 mM ATP in the pipette, holding potential (HP) = 0 mV and testing potential (TP) = -100 mV] in freshly isolated rat mesenteric smooth muscle cells showed small glibenclamide-sensitive K(ATP) currents (19.0 +/- 7.9 pA, n = 5) that increased 6.9-fold on exposure to 5 microM 14,15-EET (132.0 +/- 29.0 pA, n = 7, P < 0.05 vs. control). With 1 mM ATP in the pipette solution, K(ATP) currents (HP = 0 mV and TP = -100 mV) were increased 3.5-fold on exposure to 1 microM 14,15-EET (57.5 +/- 14.3 pA, n = 9, P < 0.05 vs. baseline). In the presence of 100 nM iberiotoxin, 1 microM 14,15-EET hyperpolarized the membrane potential from -20.5 +/- 0.9 mV at baseline to -27.1 +/- 3.0 mV (n = 6 for both, P < 0.05 vs. baseline), and the EET effects were significantly reversed by 10 microM glibenclamide (-21.8 +/- 1.4 mV, n = 6, P < 0.05 vs. EET). Incubation with 5 microM 14,15-epoxyeicosa-5(Z)-enoic acid (14,15-EEZE), a 14,15-EET antagonist, abolished the 14,15-EET effects (31.0 +/- 11.8 pA, n = 5, P < 0.05 vs. 14,15-EET, P = not significant vs. control). The 14,15-EET effects were inhibited by inclusion of anti-G(s)alpha antibody (1:500 dilution) but not by control IgG in the pipette solution. The effects of 14,15-EET were mimicked by cholera toxin (100 ng/ml), an exogenous ADP-ribosyltransferase. Treatment with the ADP-ribosyltransferase inhibitors 3-aminobenzamide (1 mM) or m-iodobenzylguanidine (100 microM) abrogated the effects of 14,15-EET on K(ATP) currents. These results were corroborated by vasodilation studies. 14,15-EET dose-dependently dilated isolated small mesenteric arteries, and this was significantly attenuated by treatment with 14,15-EEZE or 3-aminobenzamide. These results suggest that 14,15-EET activates vascular K(ATP) channels through ADP-ribosylation of G(s)alpha.  相似文献   

13.
Interactions of trimeric calf spleen purine nucleoside phosphorylase (PNP) with guanine (Gua) and its analogue, 9-deazaguanine (9-deaza-Gua), were studied by means of the steady-state fluorescence. The aim was to test the hypothesis that the enzyme stabilizes the anionic form of purine, inferred previously from the unusual increase of fluorescence observed after binding of guanine by calf spleen PNP. We have found that the dissociation constants obtained form titration experiments are in fact pH-independent in the range 7.0-10.25 for both PNP/Gua and PNP/9-deaza-Gua complexes. In particular, at pH 7.0 we found Kd = 0.12 +/- 0.02 micro M for Gua and 0.16 +/- 0.01 micro M for 9-deaza-Gua, while at the conditions where there is more than 40% of the anionic form the respective values were Kd = 0.15 +/- 0.01 micro M for Gua (pH 9.0) and 0.25 +/- 0.02 micro M for 9-deaza-Gua (pH 10.25). Hence, the enzyme does not prefer binding of anionic forms of these ligands in respect to the neutral ones. This result questions the involvement of the anionic forms in the reaction catalyzed by trimeric PNPs, and contradicts the hypothesis of a strong hydrogen bond formation between the enzyme Asn 243 residue and the purine N7 position.  相似文献   

14.
G D Lawrence  J T Spence 《Biochemistry》1975,14(16):3626-3630
The reduction of ferricytochrome c by two molybdenum(V)-cysteine complexes has been investigated as a model for electron transfer in the molybdenum enzymes sulfite oxidase and nitrate reductase. The reduction by the dioxo-bridged Mo(V)-cysteine complex, di-mu-oxo-bis-[oxo(L-cysteinato)molybdate(V)] (I), is relatively slow and its rate is first order in cyt cIII and zero order in I (k = (1.09 +/- 0.10) times 10(-3) sec minus 1, pH 7.5, 20 degrees). The reduction by the monoxo-bridged complex, mu-oxo-bis[oxodihydroxo(L-cysteinato)molybdate(V)] (II), is extremely rapid and its rate is first order in both reactants (k = (2.6 +/- 0.7) times 10(7) M minus 1 sec minus 1, pH 7.0, 25 degrees). Above pH 7.5, the reduction by II follows biphasic kinetics due to the fast reduction of a low pH form of cyt cIII and a slower reduction of a high pH form (at pH 10.0, 25 degrees, k = 2.9 times 10(6) M minus 1 sec minus 1 for the low pH form and k = 7.2 times 10(4) M minus 1 sec minus 1 for the high pH form). Reaction mechanisms for reductions by both I and II are proposed and the biological implications of the results, both for sulfite oxidase and mechanisms of electron transfer to cytochrome c, are discussed.  相似文献   

15.
Since both aminopeptidases and angiotensin I-converting enzyme are reported to degrade circulating enkephalins, we have examined the degradation of low-molecular-weight opioid peptides by a vascular plasma membrane-enriched fraction previously shown to contain both angiotensin I-converting enzyme (EC 3.4.15.1) and aminopeptidase M (EC 3.4.11.2). Except for an enkephalin analog resistant to amino-terminal hydrolysis, [D-Ala2]enkephalin, the purified vascular plasma membrane preferentially degraded low-molecular-weight opioids by hydrolysis of the N-terminal Tyr-1--Gly-2 bond. Enkephalin degradation was optimal at pH 7.0 and was inhibited by the aminopeptidase inhibitors amastatin (I50 = 0.08 microM), bestatin (9.0 microM) and puromycin (80 microM). Maximal rates of hydrolysis, calculated per mg plasma membrane protein, were highest for the shorter peptides (18.3, 15.6 and 16.6 nmol/min per mg for Met5-enkephalin, Leu5-enkephalin and Leu5-enkephalin-Arg6, respectively) and decreased with increasing peptide length (0.7 nmol/min per mg for dynorphin (1-13)). No significant hydrolysis of beta- and gamma-endorphin was detected. Km values decreased significantly with increasing peptide length (Km = 72.9 +/- 2.7, 43.6 +/- 4.7 and 21.4 +/- 0.9 microM for Met5-enkephalin, Leu5-enkephalin-Arg6 and Met5-enkephalin-Arg6-Phe7, respectively). However, no further decreases were seen with even larger sequences, i.e., dynorphin(1-13). Other peptides hydrolyzed by the plasma membrane aminopeptidase (angiotensin III, kallidin and hepta(5-11)-substance P) inhibited enkephalin degradation in a competitive manner. Thus, localization, specificity and kinetic data are consistent with identification of aminopeptidase M as a vascular enzyme with the capacity to differentially metabolize low-molecular-weight opioid peptides within the microenvironment of vascular cell surface receptors. Such differential metabolism may play a role in modulating the vascular effects of peripheral opioids.  相似文献   

16.
Andrew CR  George SJ  Lawson DM  Eady RR 《Biochemistry》2002,41(7):2353-2360
The 5-coordinate ferrous heme of Alcaligenes xylosoxidans cytochrome c' reacts with NO to form a 6-coordinate nitrosyl intermediate (lambdaSoret at 415 nm) which subsequently converts to a 5-coordinate nitrosyl end product (lambdaSoret at 395 nm) in a rate-determining step. Stopped-flow measurements at pH 8.9, 25 degrees C, yield a rate constant for the formation of the 6-coordinate nitrosyl adduct, k(on) = (4.4 +/- 0.5) x 10(4) M(-1) x s(-1), which is 3-4 orders of magnitude lower than the values for other pentacoordinate ferrous hemes and is consistent with NO binding within the sterically crowded distal heme pocket. Resonance Raman measurements of the freeze-trapped 6-coordinate nitrosyl intermediate reveal an unusually high Fe-NO stretching frequency of 579 cm(-1), suggesting a distorted Fe-N-O coordination geometry. The rate of 6- to 5-coordinate heme nitrosyl conversion is also dependent upon NO concentration, with a rate constant, k(6-5) = (8.1 +/- 0.7) x 10(3) M(-1) x s(-1), implying that an additional molecule of NO is required to form the 5c-NO adduct. Since crystallographic studies have shown that the 5-coordinate nitrosyl complex of cytochrome c' binds NO to the proximal (rather than distal) face of the heme, the NO dependence of the 6- to 5-coordinate NO conversion supports a mechanism in which the weakened His ligand, as well as the distally bound NO, is displaced by a second NO molecule which attacks and is retained in the proximal coordination position. The fact that a dependent 6- to 5-coordinate nitrosyl conversion has been previously reported for soluble guanylate cyclase suggests that the mechanism of Fe-His bond cleavage may be similar to that of cytochrome c' and strengthens the recent proposal that both proteins exhibit proximal NO binding in their 5-coordinate nitrosyl adducts.  相似文献   

17.
The physical and chemical reaction pathways of the metabolite model compound benzo(a)pyrene-7,8-diol-9,10-epoxide (BPDE) in aqueous (double-stranded) DNA solutions was investigated as a function of temperature (0-30 degrees C), pH (7.0-9.5), sodium chloride concentration (0-1.5M) and DNA concentration in order to clarify the relationships between the multiple reaction mechanisms of this diol epoxide in the presence of nucleic acids. The reaction pathways are (1) noncovalent intercalative complex formation with DNA, characterized by the equilibrium constant K, and Xb the fraction of molecules physically bound; (2) accelerated hydrolysis of BPDE bound to DNA; (3) covalent binding to DNA; and (4) hydrolysis of free BPDE(kh). The DNA-induced hydrolysis of BPDE to tetraols and the covalent binding to DNA are parallel pseudo-first-order reactions. Following the rapid (millisecond time scale) noncovalent complex formation between BPDE and DNA, a much slower (approximately minutes) H+-dependent (either specific or general acid catalysis) formation of a DNA-bound triol carbonium ion (rate constant k3) occurs. At pH 7.0 the activation energy of k3 is 8.7 +/- 0.9 kcal/mol, which is lower than the activation energy of hydrolysis of free BPDE in buffer solution (14.2 +/- 0.7 kcal/mol), and which thus partially accounts for the acceleration of hydrolysis of BPDE upon complexation with DNA. The formation of the triol carbonium ion is followed by a rapid reaction with either water to form tetraols (rate constant kT), or covalent binding to DNA (kc). The fraction of BPDE molecules which undergo covalent binding is fcov approximately equal to kc/(kc + kT) = 0.10 and is independent of the overall BPDE reaction rate constant k = kh(1 - Xb) + k3Xb if Xb----1.0, or is independent of Xb as long as k3Xb much greater than kh(1 - Xb). Thus, at Xb = 0.9, fcov is independent of pH (7.0-9.5) even though k exhibits a 70-fold variation in this pH range and k----kh above pH 9 (k3 = kh). Similarly, fcov is independent of temperature (0-30 degrees C), while k varies by a factor of approx. 3. In the range of 0-1.5 M NaCl, fcov decreases from 0.10 to 0.04. These variations are attributed to a combination of salt-induced variations in the factors k3, Xb and the ratio kc/kT.  相似文献   

18.
Bifunctional catalase-peroxidases are the least understood type of peroxidases. A high-level expression in Escherichia coli of a fully active recombinant form of a catalase-peroxidase (KatG) from the cyanobacterium Anacystis nidulans (Synechococcus PCC 6301) is reported. Since both physical and kinetic characterization revealed its identity with the wild-type protein, the large quantities of recombinant KatG allowed the examination of both the spectral characteristics and the reactivity of its redox intermediates by using the multi-mixing stopped-flow technique. The homodimeric acidic protein (pI = 4.6) contained high catalase activity (apparent K(m) = 4.8 mM and apparent k(cat) = 8850 s(-1)). Cyanide is shown to be an effective inhibitor of the catalase reaction. The second-order rate constant for cyanide binding to the ferric protein is (6.9 +/- 0.2) x 10(5) M(-1 )s(-1) at pH 7.0 and 15 degrees C and the dissociation constant of the cyanide complex is 17 microM. Because of the overwhelming catalase activity, peroxoacetic acid has been used for compound I formation. The apparent second-order rate constant for formation of compound I from the ferric enzyme and peroxoacetic acid is (1.3 +/- 0.3) x 10(4 )M(-1 )s(-1) at pH 7.0 and 15 degrees C. The spectrum of compound I is characterized by about 40% hypochromicity, a Soret region at 406 nm, and isosbestic points between the native enzyme and compound I at 355 and 428 nm. Rate constants for reduction of KatG compound I by o-dianisidine, pyrogallol, aniline and isoniazid are shown to be (7.3 +/- 0.4) x 10(6) M(-1 )s(-1), (5.4 +/- 0.3) x 10(5) M(-1 )s(-1), (1.6 +/- 0.3) x 10(5) M(-1 )s(-1) and (4.3 +/- 0.2) x 10(4) M(-1 )s(-1), respectively. The redox intermediate formed upon reduction of compound I did not exhibit the classical red-shifted peroxidase compound II spectrum which characterizes the presence of a ferryl oxygen species. Its spectral features indicate that the single oxidizing equivalent in KatG compound II is contained on an amino acid which is not electronically coupled to the heme.  相似文献   

19.
Recombinant rabbit muscle creatine kinase (CK) was titrated with MgADP in 50 mM Bicine and 5 mM Mg(OAc)2, pH 8.3, at 30.0 degrees C by following a decrease in the protein's intrinsic fluorescence. In the presence of 50 mM NaOAc, but in the absence of added creatine or nitrate, MgADP has an apparent K(d) of 135 +/- 7 microM, and the total change in fluorescence on saturation (Delta%F) is 15.3 +/- 0.3%. Acetate was used as the anion in this experiment because it does not promote the formation of a CK.MgADP.anion.creatine transition-state analogue complex (TSAC) [Millner-White and Watts (1971) Biochem. J. 122, 727-740]. In the presence of 80 mM creatine, but no nitrate, the apparent K(d) for MgADP remains essentially unchanged at 132 +/- 10 microM, while Delta%F decreases slightly to 13.2 +/- 0.3%. In the presence of 10 mM nitrate, but no creatine, the apparent K(d) is once again essentially unchanged at 143 +/- 23 microM, but the Delta%F is markedly reduced to 4.2 +/- 0.2%. The presence of both 10 mM nitrate and 80 mM creatine during titration reduces the apparent K(d) for MgADP 10-fold to 13.7 +/- 0.7 microM, and Delta%F increases to 20.6 +/- 0.3%, strongly suggesting that the simultaneous presence of saturating levels of creatine and nitrate increases the affinity of CK for MgADP and promotes the formation of the enzyme*MgADP*nitrate*creatine TSAC. When the fluorescence of CK was titrated with MgADP in the presence of 80 mM creatine and fixed saturating concentrations of various anions, apparent K(d) values for MgADP of 132 +/- 10 microM, 25.2 +/- 1.3 microM, 18.8 +/- 0.9 microM, 13.7 +/- 0.7 microM, and 6.4 +/- 0.7 microM were observed as the anion was changed from acetate to formate to chloride to nitrate to nitrite, respectively. This is the same trend reported by Millner-White and Watts for the effectiveness of various monovalent anions in forming the CK.MgADP.anion.creatine TSAC. On titration of CK with MgADP in the presence of 80 mM creatine and various fixed concentrations of NaNO3, the apparent K(d) for MgADP decreases with increasing fixed concentrations of nitrate. A plot of the apparent K(d) for MgADP vs [NO3-] suggests a K(d) for nitrate from the TSAC of 0.39 +/- 0.07 mM. Similarly, titration with MgADP in the presence of 10 mM NaNO3 and various fixed concentrations of creatine gives a value of 0.9 +/- 0.4 mM for the dissociation of creatine from the TSAC. The data were used to calculate K(TDAC), the dissociation constant of the quaternary TSAC into its individual components, of 3 x 10(-10) M3. To our knowledge this is the first reported dissociation constant for a ternary or quaternary TSAC.  相似文献   

20.
Resting potentials (Vm) were measured in mouse Leydig cells, using the whole-cell patch-clamp technique. In contrast to conventional microelectrode measurements, where a biphasic potential was observed, we recorded a stable Vm around -32.2 +/- 1.2 mV (mean +/- SEM, n = 159), at 25 degrees C, and an input resistance larger than 2.7 x 109 W. Although Vm is sensitive to changes in the extracellular concentrations of potassium and chloride, the relationship between Vm and these ions' concentrations cannot be described by either the Goldman-Hodgkin-Katz or the Nernst equation. Perifusing cells with potassium-free solution or 10?3 M ouabain induced a marked depolarization averaging 20.1 +/- 3.2 mV (n = 9) and 23.1 +/- 2.8 mV, (n = 7), respectively. Removal of potassium or addition of ouabain with the cell voltage-clamped at its Vm, resulted in an inwardly directed current, due to inhibition of the Na+K+ATPase. The pump current increased with temperature with a Q10 coefficient of 2.3 and had an average value of -6.5 +/- 0.4 pA (n = 21) at 25 degrees C. Vm also varied strongly with temperature, reaching values as low as -9.2 +/- 1.2 mV (n = 22) at 15 degrees C. Taking the pump current at 25 degrees C and a minimum estimate for the membrane input resistance, we can see that the Na+K+ATPase could directly contribute with 17.7 mV to the Vm of Leydig cells, which is a major fraction of the ?32.2 +/- 1.2 mV (n = 159) observed.  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