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 78 毫秒
1.
C K Vance  A F Miller 《Biochemistry》2001,40(43):13079-13087
Fe and Mn are both entrained to the same chemical reaction in apparently superimposable superoxide dismutase (SOD) proteins. However, neither Fe-substituted MnSOD nor Mn-substituted FeSOD is active. We have proposed that the two SOD proteins must apply very different redox tuning to their respective metal ions and that tuning appropriate for one metal ion results in a reduction potential (E(m)) for the other metal ion that is either too low (Fe) or too high (Mn) [Vance and Miller (1998) J. Am. Chem. Soc. 120, 461-467]. We have demonstrated that this is true for Fe-substituted MnSOD from Escherichia coli and that this metal ion-protein combination retains the ability to reduce but not oxidize superoxide. We now demonstrate that the corollary is also true: Mn-substituted FeSOD [Mn(Fe)SOD] has a very high E(m). Specifically, we have measured the E(m) of E. coli MnSOD to be 290 mV vs NHE. We have generated Mn(Fe)SOD and find that Mn is bound in an environment similar to that of the native (Mn)SOD protein. However, the E(m) is greater than 960 mV vs NHE and much higher than MnSOD's E(m) of 290 mV. We propose that the different tuning stems from different hydrogen bonding between the proteins and a molecule of solvent that is coordinated to the metal ion in both cases. Because a proton is taken up by SOD upon reduction, the protein can exert very strong control over the E(m), by modulating the degree to which coordinated solvent is protonated, in both oxidation states. Thus, coordinated solvent molecules may have widespread significance as "adapters" by which proteins can control the reactivity of bound metal ions.  相似文献   

2.
Phospholipase A2 activity in sonicates and acid extracts of ejaculated, washed human sperm was measured using [1-14C] oleate-labeled autoclaved E. coli and 1-[1-14C] stearoyl-2-acyl-3-sn- glycerophosphorylethanolamine as substrates. Phospholipase A was optimally active at pH 7.5, was calcium-dependent, and exclusively catalyzed the release of fatty acid from the 2-position of phospholipids. The activity was membrane-associated, and was solubilized by extraction with 0.18 N H2SO4. Acid extracts of human sperm had the highest specific activity (1709 nmols /h per mg), followed by mouse, rabbit and bull, which were 105, 36 and 1.7 nmols /h per mg, respectively. para-bromophenacyl bromide inhibited human sperm phospholipase A2 activity, but mepacrine was without effect. In the presence of 1.0 mM added CaCl2, phospholipase A2 activity was inhibited by Zn2+ and Mn2+; whereas Cu2+, Cd2+, Mg2+, or Sr2+ had no effect. Zn2+ stimulated activity at low concentrations (10(-6) to 10(-8) M), and inhibited activity in a dose-dependent manner at concentrations of 10(-5) M. The extent of stimulation by low concentrations of Zn2+ was dependent on Ca2+ concentration; at 10(-7) M, Zn2+ activity was stimulated 160% with 0.5 mM CaCl2, and only 120% with 1.0 mM CaCl2. At low concentrations (10(-5) to 10(-7) M), methoxyverapamil (D600) and trifluoperazine stimulated human sperm phospholipase A2 activity, and trifluoperazine but not D600 produced almost complete inhibition between 10(-5) and 10(-4) M of the drug. The significance of human sperm phospholipase A2 activity and its modulation by Ca2+, Zn2+ and Mn2+ in the sperm acrosome reaction is discussed.  相似文献   

3.
Martin BL  Jurado LA  Hengge AC 《Biochemistry》1999,38(11):3386-3392
Activation of calcineurin by Mn2+ and Mg2+ was compared using a heavy atom isotope analogue of the substrate p-nitrophenyl phosphate (pNPP). Heavy atom isotope effects were measured for Mg2+ activation and compared to published results of the isotope effects with Mn2+ as the activating metal. Isotope effects were measured for the kinetic parameter Vmax/Km at the nonbridging oxygen atoms [18(V/K)nonbridge]; at the position of bond cleavage in the bridging oxygen atom [18(V/K)bridge]; and at the nitrogen atom in the nitrophenol leaving group [15(V/K)]. The isotope effects increased in magnitude upon changing from an optimal pH to a nonoptimal pH; the 18(V/K)bridge effect increased from 1.0154 (+/-0.0007) to 1.0198 (+/-0.0002), and the 15(V/K) effect increased from 1.0018 (+/-0. 0002) to 1.0021 (+/-0.0003). The value for 18(V/K)nonbridge is 0. 9910 (+/-0.0003) at pH 7.0. As with Mn2+, the 18(V/K)nonbridge isotope effect indicated that the dianion was the substrate for catalysis, and that a dissociative transition state was operative for the phosphoryl transfer. Comparison to results for Mn2+ activation suggested that chemistry was more rate-limiting with Mg2+ than with Mn2+. Changing the activating metal concentration showed opposite trends with increasing Mg2+ increasing the commitment factor and seemingly making the chemistry less rate-limiting. The influence of viscosity was evaluated as well to gauge the role of chemistry. The activation of calcineurin-catalyzed hydrolysis of pNPP1 by Mg2+ or Mn2+ at pH 7.0 was compared in the presence of viscogens, glycerol and poly(ethylene glycol). Increasing glycerol caused different effects with the two activators. With Mn2+ as the activator, calcineurin activity showed a normal response with kcat and kcat/Km decreasing with viscosity. There was an inverse response with Mg2+ as the activator as values of kcat/Km increased with viscosity. From values of the normalized kcat/Km with Mn2+, the chemistry was found to be partially rate-limiting, consistent with previous heavy atom isotope studies (22). The effect observed for Mg2+ seems consistent with a change in the rate-limiting step for the two different metals at pH 7.0.  相似文献   

4.
The perfluoroalkyl compounds (PFCs), perfluoroalkyl sulfonates (PFXS) and perfluoroalkyl carboxylates (PFXA) are environmentally persistent and recalcitrant towards most conventional water treatment technologies. Here, we complete an in depth examination of the UV-254 nm production of aquated electrons during iodide photolysis for the reductive defluorination of six aquated perfluoroalkyl compounds (PFCs) of various headgroup and perfluorocarbon tail length. Cyclic voltammograms (CV) show that a potential of +2.0 V (vs. NHE) is required to induce PFC oxidation and -1.0 V is required to induce PFC reduction indicating that PFC reduction is the thermodynamically preferred process. However, PFCs are observed to degrade faster during UV(254 nm)/persulfate (S(2)O(8)(2-)) photolysis yielding sulfate radicals (E° = +2.4 V) as compared to UV(254 nm)/iodide (I(-)) photolysis yielding aquated electrons (E° = -2.9 V). Aquated electron scavenging by photoproduced triiodide (I(3)(-)), which achieved a steady-state concentration proportional to [PFOS](0), reduces the efficacy of the UV/iodide system towards PFC degradation. PFC photoreduction kinetics are observed to be dependent on PFC headgroup, perfluorocarbon chain length, initial PFC concentration, and iodide concentration. From 2 to 12, pH had no observable effect on PFC photoreduction kinetics, suggesting that the aquated electron was the predominant reductant with negligible contribution from the H-atom. A large number of gaseous fluorocarbon intermediates were semi-quantitatively identified and determined to account for ~25% of the initial PFOS carbon and fluorine. Reaction mechanisms that are consistent with kinetic observations are discussed.  相似文献   

5.
Generation of factor VIII cofactor activity requires divalent metal ions such as Ca2+ or Mn2+. Evaluation of cofactor reconstitution from isolated factor VIIIa subunits revealed the presence of a functional Ca2+ binding site within the A1 subunit. Isothermal titration calorimetry demonstrated at least two Ca2+ binding sites of similar affinity (K(d) = 0.74 microm) within the A1 subunit. Mutagenesis of an acidic residue-rich region in the A1 domain (residues 110-126) homologous to a putative Ca2+ binding site in factor V (Zeibdawi, A. R., and Pryzdial, E. L. (2001) J. Biol. Chem. 276, 19929-19936) and expression of B-domainless factor VIII molecules yielded reagents to probe Ca2+ and Mn2+ binding in a functional assay. Basal activity observed for wild type factor VIII in a metal ion-free buffer was enhanced approximately 2-fold with saturating Ca2+ or Mn2+ and yielded functional K(d) values of 1.2 and 1.40 microm, respectively. Ca2+ binding affinity was greatly reduced (or lost) in several mutants including E110A, E110D, D116A, E122A, D125A, and D126A. Alternatively, E113A, D115A, and E124A showed wild type-like activity with little or no reduction in Ca2+ affinity. However, Mn2+ affinity was minimally altered except for mutant D125A (and D116A). These results are consistent with region 110-126 serving a critical role for Ca2+ coordination with selected residues capable of contributing to a partially overlapping site for Mn2+, and that occupancy of either site is required for maximal cofactor activity.  相似文献   

6.
Although the addition of various divalent metals to beta-galactosidase resulted in apparent activation, only Mg2+ and Mn2+ actually did activate. The apparent activation by the other divalent metals was shown to be due to Mg2+ impurities. Calcium did not activate, but experiments suggested that it did bind. Other divalent metals which were studied failed to bind. The dissociation constants for Mg2+ and Mn2+ were 2.8 X 10(-7) and 1.1 X 10(-8) M, respectively, and in each case one ion bound per monomer. These constants corresponded very closely to apparent values which were obtained from activation studies. The apparent binding constant for Ca2+, obtained from competition studies, was 1.5 X 10(-5) M. Data were obtained which showed that Mg2+, Mn2+, and Ca2+ all compete for binding at a single site. Of interest and of possible molecular biological importance was the observation that, while Mg2+ bound noncooperatively (n = 1.0), Mn2+ did so in a highly cooperative manner (n = 3.4). The binding of Mn2+ (as compared to Mg2+) resulted in a twofold drop in the Vmax for the hydrolysis and transgalactosylis reactions of lactose but had little effect on the Vmax of hydrolysis of allolactose, p-nitrophenyl beta-D-galactopyranoside (PNPG), or o-nitrophenyl beta-D-galactopyranoside (ONPG); Km values were not effected differently for any of the substrates by Mn2+ as compared to Mg2+. When very low levels of divalent metal ions were present (0.01 M EDTA added) or when Ca2+ was bound with lactose as the substrate, a greater decrease was observed in the rate of the transgalactosylic reaction than in the rate of the hydrolytic reaction, and the Km values for lactose and ONPG were increased. Of the three divalent metal ions which bound to beta-galactosidase, only Mn2+ had significant stabilizing effects toward denaturing urea and heat conditions.  相似文献   

7.
Lanthanide luminescence was used to examine the effects of posttranslational adenylylation on the metal binding sites of Escherichia coli glutamine synthetase (GS). These studies revealed the presence of two lanthanide ion binding sites of GS of either adenylylation extrema. Individual emission decay lifetimes were obtained in both H2O and D2O solvent systems, allowing for the determination of the number of water molecules coordinated to each bound Eu3+. The results indicate that there are 4.3 +/- 0.5 and 4.6 +/- 0.5 water molecules coordinated to Eu3+ bound to the n1 site of unadenylylated enzyme, GS0, and fully adenylylated enzyme, GS12, respectively, and that there are 2.6 +/- 0.5 water molecules coordinated to Eu3+ at site n2 for both GS0 and GS12. Energy transfer measurements between the lanthanide donor-acceptor pair Eu3+ and Nd3+, obtained an intermetal distance measurement of 12.1 +/- 1.5 A. Distances between a Tb3+ ion at site n2 and tryptophan residues were also performed with the use of single-tryptophan mutant forms of E. coli GS. The dissociation constant for lanthanide ion binding to site n1 was observed to decrease from Kd = 0.35 +/- 0.09 microM for GS0 to Kd = 0.06 +/- 0.02 microM for GS12. The dissociation constant for lanthanide ion binding to site n2 remained unchanged as a function of adenylylation state; Kd = 3.8 +/- 0.9 microM and Kd = 2.6 +/- 0.7 microM for GS0 and GS12, respectively. Competition experiments indicate that Mn2+ affinity at site n1 decreases as a function of increasing adenylylation state, from Kd = 0.05 +/- 0.02 microM for GS0 to Kd = 0.35 +/- 0.09 microM for GS12. Mn2+ affinity at site n2 remains unchanged (Kd = 5.3 +/- 1.3 microM for GS0 and Kd = 4.0 +/- 1.0 microM for GS12). The observed divalent metal ion affinities, which are affected by the adenylylation state, agrees with other steady-state substrate experiments (Abell LM, Villafranca JJ, 1991, Biochemistry 30:1413-1418), supporting the hypothesis that adenylylation regulates GS by altering substrate and metal ion affinities.  相似文献   

8.
Further to a linear free-energy correlation of cross-reaction rate constants k12 for the reaction of eight organic radicals (OR), e.g. MV*+, from methyl viologen, with cytochrome c(III), we consider here similar studies for the reduction of the R2 protein of Escherichia coli ribonucleotide reductase, which has FeIII2 and Tyr* redox components. The same two techniques of pulse radiolysis and stopped-flow were used. Cross-reaction rate constants (22 degrees C) at pH 7.0, I=0.100 M (NaCl), were determined for the reduction of active-R2 with the eight ORs, reduction potentials E0(1) from -0.446 to +0.194 V. Samples of active-R2 have an FeIII2 met-R2 component, which in the present studies was close to 40%. Concurrent reactions have to be taken into account for the five most reactive ORs, corresponding to reduction of the FeIII2 of met-R2 and then of active-R2. Separate experiments on met-R2 reproduced the first of these rate constants, which on average is approximately 66% larger than the second rate constant. A single Marcus free-energy plot of log k12-0.5 log10f versus -E0(1)/0.059 describes all the data and the slope of 0.54 is in satisfactory agreement with the theoretical value of 0.50. Such behaviour is unexpected since the Tyr* is a much stronger oxidant (E0 approximately 1.0 V versus NHE) as compared to FeIII2 (E0 close to zero). X-ray structures of the met- and red-R2 states have indicated that electroneutrality of the approximately 10 A buried active site is maintained. Proton transfer is therefore proposed as a rapid sequel to electron transfer. Other reactions considered are the much slower conventional time-range reductions of active-R2 with hydrazine and dithionite. For these reactions one and/or two-equivalent changes are possible. With both reductants, met-R2 reacts about four-fold faster than active-R2, and as with the ORs the less strongly oxidising FeIII2 component is reduced before the Tyr*.  相似文献   

9.
Hwang HJ  McLain A  Debus RJ  Burnap RL 《Biochemistry》2007,46(47):13648-13657
The light-driven, oxidative assembly of Mn2+ ions into the H2O-oxidation complex (WOC) of the photosystem II (PSII) reaction center is termed photoactivation and culminates in the formation of the oxygen-evolving (Mn4-Ca) center of the WOC. Initial binding and photooxidation of Mn2+ to the apoprotein is critically dependent upon aspartate 170 of the D1 protein (D1-D170) of the high affinity Mn site [Nixon and Diner (1992) Biochemistry 31, 942-948]. Three O2-evolving mutant strains of Synechocystis, D1-D170E, D1-D170H, and D1-D170V, were studied in terms of the kinetics of photoactivation under both continuous and flashing light. Photoactivation using single turnover flashes revealed D1-D170H and D1-D170V, but not D1-D170E, were prone to form substantial amounts ( approximately 40-50%) of inactive centers ascribed to photoligation of aberrant nonfunctional Mn based upon the reversibility of the inactivation and similarity to previous in vitro results [Chen, C., Kazimir, J., and Cheniae, G. M. (1995) Biochemistry 34, 13511-13526]. On the other hand, D1-D170E lowers the quantum efficiency of photoactivation compared to the wild-type by the largest amount (80% decrease) versus D1-D170H and D1-D170V, which do not produce measurable decreases in quantum efficiency. The low quantum efficiency of photoactivation in D1-D170E is due to the destabilization of photoactivation intermediates. Numerical analysis indicates that the PSII centers in D1-D170E are heterogeneous with respect to photoactivation kinetics and that the majority of centers are characterized by intermediates that decay approximately 10-fold more rapidly than the wild-type control. Additionally, the kinetics of O2 release during the S3-S0 transition was markedly retarded in D1-D170E, in contrast to D1-D170H and D1-D170V, which did not exhibit a discernible slow-down compared to the wild-type.  相似文献   

10.
Smooth muscle membrane potential (E(m)) depends on K(+) channels, and arteries from rats made hypertensive with N(omega)-nitro-l-arginine (LHR) are depolarized compared with control. We hypothesized that decreased K(+) channel function, due to decreased K(+) channel protein expression, underlies E(m) depolarization. Furthermore, K(+) channel blockers should move control E(m) (-46 +/- 1 mV) toward that in LHR (-37 +/- 2 mV) and normalize contraction. The E(m) vs. K(+) relationship was less steep in LHR (23 +/- 2 vs. 28 +/- 1 mV/log K(+) concentration), and contractile sensitivity to K(+) was increased (EC(50) = 37 +/- 1 vs. 23 +/- 1 mM). Iberiotoxin (10 nM), an inhibitor of large-conductance Ca(2+)-activated K(+) (BK(Ca)) channels, depolarized control and LHR E(m) to -35 +/- 1 and -30 +/- 2 mV, respectively; however, effects on K(+) sensitivity were more profound in LHR (EC(50) = 25 +/- 2 vs. 15 +/- 3 mM). The voltage-dependent K(+) (K(V)) channel blocker 4-aminopyridine (3 mM) depolarized control E(m) to the level of LHR (-28 +/- 1 vs. -28 +/- 1 mV); however, effects on K(+) sensitivity were greater in LHR (EC(50) = 17 +/- 4 vs. 4 +/- 4 mM). Western blots revealed reduced BK(Ca) and K(V)1.5 channel expression in LHR arteries. The findings suggest that diminished expression of K(+) channels contributes to depolarization and enhanced contractile sensitivity. These conclusions are supported by direct electrophysiological assessment of BK(Ca) and K(V) channel function in control and LHR smooth muscle cells.  相似文献   

11.
This study explored mitochondrial capacities to oxidize carbohydrate and fatty acids and functional optimization of mitochondrial respiratory chain complexes in athletes who regularly train at high exercise intensity (ATH, n = 7) compared with sedentary (SED, n = 7). Peak O(2) uptake (Vo(2max)) was measured, and muscle biopsies of vastus lateralis were collected. Maximal O(2) uptake of saponin-skinned myofibers was evaluated with several metabolic substrates [glutamate-malate (V(GM)), pyruvate (V(Pyr)), palmitoyl carnitine (V(PC))], and the activity of the mitochondrial respiratory complexes II and IV were assessed using succinate (V(s)) and N,N,N',N'-tetramethyl-p-phenylenediamine dihydrochloride (V(TMPD)), respectively. Vo(2max) was higher in ATH than in SED (57.8 +/- 2.2 vs. 31.4 +/- 1.3 ml.min(-1).kg(-1), P < 0.001). V(GM) was higher in ATH than in SED (8.6 +/- 0.5 vs. 3.3 +/- 0.3 micromol O(2).min(-1).g dry wt(-1), P < 0.001). V(Pyr) was higher in ATH than in SED (8.7 +/- 1.0 vs. 5.5 +/- 0.2 micromol O(2).min(-1).g dry wt(-1), P < 0.05), whereas V(PC) was not significantly different (5.3 +/- 0.9 vs. 4.4 +/- 0.5 micromol O(2).min(-1).g dry wt(-1)). V(S) was higher in ATH than in SED (11.0 +/- 0.6 vs. 6.0 +/- 0.3 micromol O(2).min(-1).g dry wt(-1), P < 0.001), as well as V(TMPD) (20.1 +/- 1.0 vs. 16.2 +/- 3.4 micromol O(2).min(-1).g dry wt(-1), P < 0.05). The ratios V(S)/V(GM) (1.3 +/- 0.1 vs. 2.0 +/- 0.1, P < 0.001) and V(TMPD)/V(GM) (2.4 +/- 1.0 vs. 5.2 +/- 1.8, P < 0.01) were lower in ATH than in SED. In conclusion, comparison of ATH vs. SED subjects suggests that regular endurance training at high intensity promotes the enhancement of maximal mitochondrial capacities to oxidize carbohydrate rather than fatty acid and induce specific adaptations of the mitochondrial respiratory chain at the level of complex I.  相似文献   

12.
Legler PM  Lee HC  Peisach J  Mildvan AS 《Biochemistry》2002,41(14):4655-4668
Escherichia coli GDP-mannose mannosyl hydrolase (GDPMH), a homodimer, catalyzes the hydrolysis of GDP-alpha-D-sugars to yield the beta-D-sugar and GDP by nucleophilic substitution with inversion at the C1' carbon of the sugar [Legler, P. M., Massiah, M. A., Bessman, M. J., and Mildvan, A. S. (2000) Biochemistry 39, 8603-8608]. GDPMH requires a divalent cation for activity such as Mn2+ or Mg2+, which yield similar kcat values of 0.15 and 0.13 s(-1), respectively, at 22 degrees C and pH 7.5. Kinetic analysis of the Mn2+-activated enzyme yielded a K(m) of free Mn2+ of 3.9 +/- 1.3 mM when extrapolated to zero substrate concentration (K(a)Mn2+), which tightened to 0.32 +/- 0.18 mM when extrapolated to infinite substrate concentration (K(m)Mn2+). Similarly, the K(m) of the substrate extrapolated to zero Mn2+ concentration (K(S)(GDPmann) = 1.9 +/- 0.5 mM) and to infinite Mn2+ concentration (K(m)(GDPmann) = 0.16 +/- 0.09 mM) showed an order of magnitude decrease at saturating Mn2+. Such mutual tightening of metal and substrate binding suggests the formation of an enzyme-metal-substrate bridge complex. Direct Mn2+ binding studies, monitoring the concentration of free Mn2+ by EPR and of bound Mn2+ by its enhanced paramagnetic effect on the longitudinal relaxation rate of water protons (PRR), detected three Mn2+ binding sites per enzyme monomer with an average dissociation constant (K(D)) of 3.2 +/- 1.0 mM, in agreement with the kinetically determined K(a)Mn2+. The enhancement factor (epsilon(b)) of 11.5 +/- 1.2 indicates solvent access to the enzyme-bound Mn2+ ions. No cross relaxation was detected among the three bound Mn2+ ions, suggesting them to be separated by at least 10 A. Such studies also yielded a weak dissociation constant for the binary Mn2+-GDP-mannose complex (K1 = 6.5 +/- 1.0 mM) which significantly exceeded the kinetically determined K(m) values of Mn2+, indicating the true substrate to be GDP-mannose rather than its Mn2+ complex. Substrate binding monitored by changes in 1H-15N HSQC spectra yielded a dissociation constant for the binary E-GDP-mannose complex (K(S)(GDPmann)) of 4.0 +/- 0.5 mM, comparable to the kinetically determined K(S) value (1.9 +/- 0.5 mM). To clarify the metal stoichiometry at the active site, product inhibition by GDP, a potent competitive inhibitor (K(I) = 46 +/- 27 microM), was studied. Binding studies revealed a weak, binary E-GDP complex (K(D)(GDP) = 9.4 +/- 3.2 mM) which tightened approximately 500-fold in the presence of Mn2+ to yield a ternary E-Mn2+-GDP complex with a dissociation constant, K3(GDP) = 18 +/- 9 microM, which overlaps with the K(I)(GDP). The tight binding of Mn2+ to 0.7 +/- 0.2 site per enzyme subunit in the ternary E-Mn2+-GDP complex (K(A)' = 15 microM) and the tight binding of GDP to 0.8 +/- 0.1 site per enzyme subunit in the ternary E-Mg2+-GDP complex (K3 < 0.5 mM) indicate a stoichiometry close to 1:1:1 at the active site. The decrease in the enhancement factor of the ternary E-Mn2+-GDP complex (epsilon(T) = 4.9 +/- 0.4) indicates decreased solvent access to the active site Mn2+, consistent with an E-Mn2+-GDP bridge complex. Fermi contact splitting (4.3 +/- 0.2 MHz) of the phosphorus signal in the ESEEM spectrum established the formation of an inner sphere E-Mn2+-GDP complex. The number of water molecules coordinated to Mn2+ in this ternary complex was determined by ESEEM studies in D2O to be two fewer than on the average Mn2+ in the binary E-Mn2+ complexes, consistent with bidentate coordination of enzyme-bound Mn2+ by GDP. Kinetic, metal binding, and GDP binding studies with Mg2+ yielded dissociation constants similar to those found with Mn2+. Hence, GDPMH requires one divalent cation per active site to promote catalysis by facilitating the departure of the GDP leaving group, unlike its homologues the MutT pyrophosphohydrolase, which requires two, or Ap4A pyrophosphatase, which requires three.  相似文献   

13.
S G Rhee  P B Chock 《Biochemistry》1976,15(8):1755-1760
The kinetics of protein fluorescence change exhibited by ADP or orthophosphate addition to the Mg2+-or Mn2+-activated unadenylylated glutamine synthetase from Escherichia coli were studied. The kinetic patterns of these reactions are incompatible with a simple bimolecular binding process and a mechanism which required protein isomerization prior to substrate binding. They are consistent with a mechanism in which direct substrate binding is followed by a substrate-induced conformational change step, ES in equilibrium ES. At pH 7.0 and 15 degrees C, the association constants for the direct binding (K1) of ADP to MnE1.0 and of Pi to MnE1.0ADP are 3.9 X 10(4) and 2.28 X 10(2) M(-1), respectively. The association constant for the direct binding of ADP to MnE1.0Pi is 2.3 X 10(4) M(-1) at pH 7.0 and 19 degrees C. The deltaG degrees for the substrate-induced conformational step are -3.5 and -1.3 kcal mol(-1) due to ADP binding to MnE1.0Pi and MnE1.0, respectively, and -1.4 kcal mol(-1) due to Pi binding to MnE1.0ADP. Rate constants, k2 and k(-2), for the isomerization step are: 90 and 9.5 s(-1) for ADP binding to MnE1.0, 440 and 0.36 s(-1) for ADP binding to MnE1.0Pi, and 216 and 1.8 s(-1) for Pi binding to MnE1.0ADP. Due to low substrate affinity, the association constant for direct Pi binding to MnE1.0 was roughly estimated to be 230 M(-1) and k2 = 750 s(-1), k(-2) = 250 s(-1). At 9 degrees C and pH 7.0, the estimated association constants for the direct ADP binding to MgE1.0 and MgE1.0 Pi are 1.8 X 10(4) and 1.6 X 10(4) M(-1), respectively; and the rate constants for the isomerization step associated with the corresponding reaction are k2 = 550 s(-1), k(-2) = 500 s(-1), and k2 = 210 s(-1), k(-2) = 100 s(-1). From the kinetic analysis it is evident that the inability of Mn2+ to support biosynthetic activity of the unadenylylated enzyme is due to the slow rate of ADP release from the MnE1.0PiADP complex. In contrast the large k(-2) obtained for ADP release from the MgE1.0ADP or MgE1.0PiADP complex indicates that this step is not rate limiting in the biosynthesis of glutamine since the k catalysis obtained under the same conditions is 7.2 s(-1).  相似文献   

14.
The purpose of this study was to examine ethnic differences in the metabolic responses to a 16-week intervention designed to improve insulin sensitivity (SI), adiposity, and inflammation in obese African-American and Latino adolescents. A total of 100 participants (African Americans: n = 48, Latino: n = 52; age: 15.4 ± 1.1 years, BMI percentile: 97.3 ± 3.3) were randomly assigned to interventions: control (C; n = 30), nutrition (N; n = 39, 1×/week focused on decreasing sugar and increasing fiber intake), or nutrition + strength training (N+ST; n = 31, 2×/week). The following were measured at pre- and postintervention: strength, dietary intake, body composition (dual-energy X-ray absorptiometry/magnetic resonance imaging) and glucose/insulin indexes (oral glucose tolerance test (OGTT)/intravenous glucose tolerance test (IVGTT)) and inflammatory markers. Overall, N compared to C and N+ST reported significant improvements in SI (+16.5% vs. -32.3% vs. -6.9% respectively, P < 0.01) and disposition index (DI: +15.5% vs. -14.2% vs. -13.7% respectively, P < 0.01). N+ST compared to C and N reported significant reductions in hepatic fat fraction (HFF: -27.3% vs. -4.3% vs. 0% respectively, P < 0.01). Compared to N, N+ST reported reductions in plasminogen activator inhibitor-1 (PAI-1) (-38.3% vs. +1.0%, P < 0.01) and resistin (-18.7% vs. +11.3%, P = 0.02). There were no intervention effects for all other measures of adiposity or inflammation. Significant intervention by ethnicity interactions were found for African Americans in the N group who reported increases in total fat mass, 2-h glucose and glucose incremental areas under the curve (IAUC) compared to Latinos (P's < 0.05). These interventions yielded differential effects with N reporting favorable improvements in SI and DI and N+ST reporting marked reductions in HFF and inflammation. Both ethnic groups had significant improvements in metabolic health; however some improvements were not seen in African Americans.  相似文献   

15.
This study examines the activation of divalent cation entry into rat parotid gland acinar cells by using Mn2+ as a Ca2+ surrogate cation. Following muscarinic-cholinergic stimulation of dispersed parotid acini with carbachol (10 microM), the onset of internal Ca2+ release (cytosolic [Ca2+], [Ca2+]i, increase) and the stimulation of Mn2+ entry (increase in fura2 quenching) are not simultaneously detected. [Ca2+]i elevation, due to intracellular release, is detected almost immediately following carbachol addition and peak [Ca2+]i increase occurs at 6.0 +/- 0.8 sec. However, there is an interval (apparent lag) between carbachol addition and the detection of stimulated Mn2+ entry. This apparent lag is decreased from 26 +/- 3.1 sec to 9.2 +/- 1.5 sec when external Mn2+ ([Mn2+]0) is increased from 12.5 to 500 microM. It is not decreased further with increase in [Mn2+]0 from 500 microM to 1 mM (9.8 +/- 2.1 sec), although both intracellular free Mn2+ and [Mn2+-fura2]/[fura2] increase. Thus, at [Mn2+]0 < 500 microM, the observed lag time is partially due to a limitation in the magnitude of Mn2+ entry. Furthermore, neither peak [Ca2+]i nor the time required to reach peak [Ca2+]i is significantly altered by [Mn2+]0 (12.5 microM to 1 mM). At every [Mn2+]0 tested (i.e., 12.5 microM-1 mM), the apparent lag is significantly greater than the time required to reach peak [Ca2+]i. However, when carbachol stimulation of the [Ca2+]i increase is attenuated by loading the acini with the Ca2+ chelator, 2-bis(O-aminophenoxy)ethane-N,N,N',N'-tetraacetate (BAPTA), there is no detectable lag in carbachol stimulation of Mn2+ entry (with 1 mM [Mn2+]0). Importantly, in BAPTA-loaded acini, carbachol stimulates Mn2+ entry via depletion of the internal Ca2+ pool and not via direct activation of other divalent cation entry mechanisms. Based on these results, we suggest that the apparent lag in the detection of carbachol stimulation of Mn2+ entry into parotid acinar cells is due to a retardation of Mn2+ entry by the initial increase in [Ca2+]i, due to internal release, which most likely occurs proximate to the site of divalent cation entry.  相似文献   

16.
S(IV) (SO(2),HSO(3)(-)andSO(3)(2-)) autoxidation catalyzed by Cu(II)/tetraglycine complexes in the presence of DNA or 2'-deoxyguanosine (dGuo) resulted in DNA strand breaks and formation of 8-oxo-7,8-dihydro-2'-deoxyguanosine (8-oxodGuo), respectively. Ni(II), Co(II) or Mn(II) (1.0x10(-4)M) complexes had much smaller effects. Cu(II)/tetraglycine (1.0x10(-4)M) in the presence of Ni(II) or Mn(II) (10(-7)-10(-6)M) and S(IV) showed remarkable synergistic effect with these metal ions producing a higher yield of 8-oxodGuo. Oxidation of dGuo and DNA damage were attributed to oxysulfur radicals formed as intermediates in S(IV) autoxidation catalyzed by transition metal ions. SO*(3)(-) and HO* radicals were detected by EPR-spin trapping experiments with DMPO (5,5-dimethyl-1-pyrroline-N-oxide).  相似文献   

17.
We examined the effect of inhibition of Na+/H+ exchange (NHE) on duodenal bicarbonate secretion (DBS) in rats to further understand DBS regulation. DBS was measured by using the pH-stat method and by using CO2-sensitive electrodes. 5-(N,N-dimethyl)-amiloride (50 microM; DMA), a concentration that selectively inhibits the NHE isoforms NHE1 and NHE2, but not NHE3, did not affect DBS. Nevertheless, 3 mM DMA, a higher concentration that inhibits NHE1, NHE2, and NHE3, significantly increased DBS. Moreover, S1611 and S3226, both specific inhibitors of NHE3 only, or perfusion with Na+-free solutions, dose dependently increased DBS, as measured by pH-stat and CO2-sensitive electrode, without affecting intracellular pH. Coperfusion with 0.1 microM indomethacin, 0.5 mM DIDS, or 1 mM methazolamide did not affect S3226-induced DBS. Nevertheless, coperfusion with 0.1 and 0.3 mM 5-nitro-2-(3-phenylpropylamino) benzoic acid, which inhibits the cystic fibrosis transmembrane conductor regulator (CFTR), dose dependently inhibited S3226-induced DBS. In conclusion, only specific apical NHE3 inhibition increased DBS, whereas prostaglandin synthesis, Na+-HCO3- cotransporter activation, or intracellular HCO3- formation by carbonic anhydrase was not involved. Because NHE3 inhibition-increased DBS was inhibited by an anion channel inhibitor and because reciprocal CFTR regulation has been previously shown between NHE3 and apical membrane anion transporters, we speculate that NHE3 inhibition increased DBS by altering anion transporter function.  相似文献   

18.
Mn porphyrins are among the most efficient SOD mimics with potency approaching that of SOD enzymes. The most potent ones, Mn(III) N-alkylpyridylporphyrins bear positive charges in a close proximity to the metal site, affording thermodynamic and kinetic facilitation for the reaction with negatively charged superoxide. The addition of electron-withdrawing bromines onto beta-pyrrolic positions dramatically improves thermodynamic facilitation for the O2*- dismutation. We have previously characterized the para isomer, Mn(II)Br(8)TM-4-PyP(4+) [Mn(II) beta-octabromo-meso-tetrakis(N-methylpyridinium-4-yl)porphyrin]. Herein we fully characterized its meta analogue, Mn(II)Br(8)TM-3-PyP(4+) with respect to UV/vis spectroscopy, electron spray mass spectrometry, electrochemistry, O2*- dismutation, metal-ligand stability, and the ability to protect SOD-deficient Escherichia coli in comparison with its para analogue. The increased electron-deficiency of the metal center stabilizes Mn in its +2 oxidation state. The metal-centered Mn(III)/Mn(II) reduction potential, E((1/2))=+468 mV vs NHE, is increased by 416 mV with respect to non-brominated analogue, Mn(III)TM-3-PyP(5+) and is only 12 mV less positive than for para isomer. Yet, the complex is significantly more stable towards the loss of metal than its para analogue. As expected, based on the structure-activity relationships, an increase in E((1/2)) results in a higher catalytic rate constant for the O2*- dismutation, log k(cat)> or =8.85; 1.5-fold increase with respect to the para isomer. The IC(50) was calculated to be < or =3.7 nM. Manipulation of the electron-deficiency of a cationic porphyrin resulted, therefore, in the highest k(cat) ever reported for a metalloporphyrin, being essentially identical to the k(cat) of superoxide dismutases (log k(cat)=8.84-9.30). The positive kinetic salt effect points to the unexpected, unique and first time recorded behavior of Mn beta-octabrominated porphyrins when compared to other Mn porphyrins studied thus far. When species of opposing charges react, the increase in ionic strength invariably results in the decreased rate constant; with brominated porphyrins the opposite was found to be true. The effect is 3.5-fold greater with meta than with para isomer, which is discussed with respect to the closer proximity of the quaternary nitrogens of the meta isomer to the metal center than that of the para isomer. The potency of Mn(II)Br(8)TM-3-PyP(4+) was corroborated by in vivo studies, where 500 nM allows SOD-deficient E. coli to grow >60% of the growth of wild type; at concentrations > or =5 microM it exhibits toxicity. Our work shows that exceptionally high k(cat) for the O2*- disproportionation can be achieved not only with an N(5)-type coordination motif, as rationalized previously for aza crown ether (cyclic polyamines) complexes, but also with a N(4)-type motif as in the Mn porphyrin case; both motifs sharing "up-down-up-down" steric arrangement.  相似文献   

19.
With aging, the kidney develops a progressive deterioration of several structures and functions. Proximal tubular acidification is impaired in old rats with a decrease in the activity of brush border Na+/H+ exchange and a fall of H-ion flux measured with micropuncture experiments. In the present work we evaluate the contribution of 5-N-ethyl-n-isopropyl amiloride- (EIPA) and bafilomycin-sensitive bicarbonate flux (JHCO3-) in proximal convoluted tubules of young and aged rats. We performed micropuncture experiments inhibiting the Na+/H+ exchanger with EIPA (10(-4) M) and the V-H+ATPase with bafilomycin (10(-6) M). We used antibodies against the NHE3 isoform of the Na+/H+ exchanger and the subunit E of the V-H+ATPase for detecting by Western blot the abundance of these proteins in brush border membrane vesicles from proximal convoluted tubules of young and old rats. The abundance of NHE3 and the V-H+ATPase was similar in 18-month-old and 3-month-old rats. The bicarbonate flux in old rats was 30% lower than in young rats. EIPA reduced by 60% and bafilomycin by 30% in young rats; in contrast, EIPA reduced by approximately 40% and bafilomycin by approximately 50% in old rats. The inhibited by bafilomycin was the same in young and old rats: 0.62 nmol.cm-2.s-1 and 0.71 nmol.cm-2.s-1, respectively. However, the EIPA-sensitive fraction was larger in young than in old rats: 1.26 nmol.cm-2.s-1 vs. 0.85 nmol.cm-2.s-1, respectively. These results suggest that the component more affected in bicarbonate reabsorption of proximal convoluted tubules from aged rats is the Na+-H+ exchanger, probably a NHE isoform different from NHE3.  相似文献   

20.
We discovered that 4-(2-pyridylazo) resorcinol (PAR) has a strong catalytic effect on luminol-potassium ferricyanide chemiluminescence (CL). Results indicated that the chemiluminescence intensities at maximum light emission were linearly corrected with the concentration of PAR over the range 1.0 x 10(-5)-1.0 x 10(-7) mol/L. A detection limit of 5.7 x 10(-8) mol/L for PAR was achieved. It was found that some metal ions strongly affected this catalytic reaction. Based on this finding, the luminol-potassium ferricyanide-PAR reaction was developed for the determination of metal ions. The detection limits (S/N = 3) for Ni2+, Cr3+, Zn2+, Co2+ and Mn2+ were determined to be 1.0 x 10(-9) mol/L, 5.0 x 10(-9) mol/L, 5.0 x 10(-8) mol/L, 1.0 x 10(-9) mol/L and 1.0 x 10(-8) mol/L, respectively. In addition, the relative standard deviation values for these metal ion assays were in the range 0.82-2.72% (n = 6).  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