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.
The effect of oxygen on the radiolysis of tyrosine in aqueous solutions was investigated by using gamma and pulsed electron irradiation. Steady-state radiolysis was reexamined and extended to include the effect of pH and determination of hydrogen peroxide. The loss of tyrosine, G(-Tyr), during irradiation and yields of 3,4-dihydroxyphenylalanine, G(DOPA), and hydrogen peroxide, G(H2O2), are determined in the pH range from 1 to 9. In the whole pH range used G(-Tyr) equals G(DOPA), and a higher G(H2O2) than expected was observed. In slightly acid and neutral media, both G(-Tyr) and G(DOPA) equal the yield of hydroxyl radicals, GOH, formed in the radiolysis of water, while the excess of hydrogen peroxide equals 1/2 GOH. Hence it was concluded that all tyrosine OH-adducts react with oxygen yielding peroxy radicals. In acid and alkaline media all measured yields decrease. This is caused by formation of tyrosine phenoxyl radicals (TyrO), which react with superoxide anion (O2-) and hydroperoxy (HO2) radicals regenerating tyrosine. By using pulse radiolysis K(TyrO + O2) less than or equal to 2 X 10(5) mol-1 dm3 s-1 and k(TyrO + O2-) = (1.7 +/- 0.2) X 10(9) mol-1 dm3 s-1 were determined. On the basis of the results, a reaction mechanism is proposed.  相似文献   

2.
The Eastern oyster, Crassostrea virginica, lives in shallow coastal waters and experiences many different environmental extremes including hypoxia, hypercapnia and air exposure and many oysters are infected with the protozoan parasite Perkinsus marinus. The effects of these conditions on oyster metabolism, as measured by oxygen uptake, were investigated. Mild hypercapnia had no effect on the ability of oysters to regulate oxygen uptake in hypoxic water, as measured by the B2 coefficient of oxygen regulation. The average B2 was -0.060x10(-3) (+/-0.01x10(-3) S.E.M.; n=20; low and high CO(2) treatments combined) in oysters uninfected with P. marinus and -0.056x10(-3) (+/-0.01x10(-3) S.E.M.; n=16; low and high CO(2) treatments combined) in infected oysters. There was no significant effect of light to moderate infections of P. marinus on oxygen regulation. Nor did the presence of P. marinus have an effect on the rate of oxygen uptake of whole animals in well-aerated water. In well-aerated conditions, oxygen uptake was significantly reduced by moderate hypercapnia in oysters when data from uninfected and infected oysters were combined. Mean oxygen uptake of infected oysters under hypercapnia (pCO(2)=6-8 Torr; pH 7) was 9.10 μmol O(2) g ww(-1) h(-1) +/-0.62 S.E.M. (n=9), significantly different from oxygen uptake under normocapnia (pCO(2) 相似文献   

3.
A method for recording O2 concentrations in nonconducting organic media with the Clark oxygen electrode was developed. Spontaneous oxidation of Na2S2O4 and the enzymatic reduction of NaBO3 or H2O2 by bovine liver catalase trapped in hydrated micelles of dioctylsulfosuccinate (AOT)/toluene were used as model systems. O2 titration with the above systems showed that air-saturated 1.6 M H2O/0.2 M AOT/toluene media contain seven times more O2 (1.4 mM) than aqueous solutions (0.2 mM). The measured Km values of catalase for NaBO3 and H2O2 in organic media were Kmov = 15 and 17 mM, respectively, whereas in aqueous buffer the values were 45 and 54 mM. In the toluene media, catalase activity increased with the W0 (H2O/AOT molar ratio) of the micellar preparation, reaching maximal activity at W0 = 10-12; under this condition, the catalytic center activity (Kp) of H2O2 was 7 x 10(6) min-1, similar to that obtained in the aqueous buffer (H2O2 = 7 x 10(6) min-1). It was found that the optimal pH for catalase in toluene media (pH 8.0) was shifted 1.0 unit compared to that in the aqueous buffer (pH 7.0). On the other hand, catalase was severely inhibited by NaN3 in both media. Thus, polarography based on the Clark oxygen electrode seems to be an easy, rapid, and sensitive technique for studying enzyme reactions consuming or evolving O2 in apolar media.  相似文献   

4.
Iron loss in vitro by the iron scavenger bovine lactoferrin was investigated in acidic media in the presence of three different monoanions (NO(3)(-), Cl(-) and Br(-)) and one dianion (SO(4)(2-)). Holo and monoferric C-site lactoferrins lose iron in acidic media (pH< or =3.5) by a four-step mechanism. The first two steps describe modifications in the conformation affecting the whole protein, which occur also with apolactoferrin. These two processes are independent of iron load and are followed by a third step consisting of the gain of two protons. This third step is kinetically controlled by the interaction with two Cl(-), Br(-) and NO(3)(-) or one SO(4)(2-). In the fourth step, iron loss is under the kinetic control of a slow gain of two protons; third-order rate-constants k(2), 4.3(+/-0.2)x10(3), 3.4(+/-0.5)x10(3), 3.3(+/-0.5)x10(3) and 1.5(+/-0.5)x10(3) M(-2) s(-1) when the protein is in interaction with SO(4)(2-), NO(3)(-), Cl(-) or Br(-), respectively. This step is accompanied by the loss of the interaction with the anions; equilibrium constant K(2), 20+/-5 mM, 1.0(+/-0.2)x10(-1), 1.5(+/-0.5)x10(-1) and 1.0(+/-0.3)x10(-1) M(2), for SO(4)(-), NO(3)(-), Cl(-) and Br(-), respectively. This mechanism is very different from that determined in mildly acidic media at low ionic strength (micro<0.5) for the iron transport proteins, serum transferrin and ovotransferrin, with which no prior change in conformation or interaction with anions is required. These differences may result from the fact that in the transport proteins, the interdomain hydrogen bonds that consolidate the closed conformation of the iron-binding cleft occur between amino acid side-chain residues that can protonate in mildly acidic media. With bovine lactoferrin, most of the interdomain hydrogen bonds involved in the C-site and one of those involved in the N-site occur between amino acid side-chain residues that cannot protonate. The breaking of the interdomain H-bond upon protonation can trigger the opening of the iron cleft, facilitating iron loss in serum transferrin and ovotransferrin. This situation is, however, different in lactoferrin, where iron loss requires a prior change in conformation. This can explain why lactoferrin does not lose its iron load in acidic media and why it is not involved in iron transport in acidic endosomes.  相似文献   

5.
The promising advantages of Prussian Blue (PB) as catalyst and of the thick film screen printing technology have been combined to assemble sensors with improved characteristics for the amperometric determination of H(2)O(2). PB-modified screen printed electrodes were applied to detect H(2)O(2) at an applied potential of -0.05 V versus the internal screen printed Ag pseudoreference electrode, showing a detection limit of 10(-7) mol l(-1), a linearity range from 10(-7) to 5x10(-5) mol l(-1), a sensitivity of 234 microA mmol l(-1) cm(-2), and a high selectivity. Improved stability at alkaline pH values was also observed, which made possible their use with enzymes having an optimum basic pH. Then, the immobilisation of a single enzyme (glucose oxidase (GOD) or choline oxidase (ChOX)) or of two enzymes, acetylcholinesterase (AchE) coimmobilised with ChOX, has been performed on the surface of PB modified screen-printed electrodes (SPEs) using glutaraldehyde and Nafion. ChOX has been selected as an example of enzyme working at alkaline pH. The choline biosensors showed a detection limit of 5x10(-7) mol l(-1), a wide linearity range (5x10(-7)-10(-4) mol l(-1)), a high selectivity and a remarkable long term stability of 9 months at 4 degrees C, and at least 4 weeks at room temperature. Similar analytical characteristics and stability were observed with the acetylcholine biosensors.  相似文献   

6.
An assay system for detection of Escherichia coli O157:H7 was developed based on immunomagnetic separation of the target pathogen from samples and absorbance measurement of p-nitrophenol at 400 nm from p-nitrophenyl phosphate hydrolysis by alkaline phosphatase (EC 3.1.3.1) on the "sandwich" structure complexes (antibodies coated onto micromagnetic beads--E. coli O157:H7-antibodies conjugated with the enzyme) formed on the microbead surface. The effects of immunoreaction time, phosphate buffer concentration, pH and temperature on the immunomagnetic separation of E. coli O157:H7 from samples were determined and the conditions used for the separation were 1-h reaction time, 1.0 x 10(-2) M PBS, pH 8.0 and 33 degrees C in this system. The effects of MgCl(2) concentration, Tris buffer concentration, pH and temperature on the activity of alkaline phosphatase conjugated on the immuno-"sandwich" structure complexes were investigated after immunomagnetic separation of the target pathogen and the conditions used for the enzymatic amplification were 1.0 x 10(-4) M MgCl(2), 1.0 M Tris buffer, pH 8.0, 28 degrees C and 30-min reaction time during the assay. The selectivity of the system was examined and no interference from the other pathogens including Salmonella typhimurium, Campylobacter jejuni and Listeria monocytogenes was observed. Its working range was from 3.2 x 10(2) to 3.2 x 10(4) CFU/ml, and the relative standard deviation was 2.5-9.9%. The total detection time was less than 2 h.  相似文献   

7.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

8.
A study is presented on proton transfer associated with the reaction of the fully reduced, purified bovine heart cytochrome c oxidase with molecular oxygen or ferricyanide. The proton consumption associated with aerobic oxidation of the four metal centers changed significantly with pH going from approximately 3.0 H(+)/COX at pH 6.2-6.3 to approximately 1.2 H(+)/COX at pH 8.0-8.5. Rereduction of the metal centers was associated with further proton uptake which increased with pH from approximately 1.0 H(+)/COX at pH 6.2-6.3 to approximately 2.8 H(+)/COX at pH 8.0-8.5. Anaerobic oxidation of the four metal centers by ferricyanide resulted in the net release of 1.3-1.6 H(+)/COX in the pH range 6.2-8.2, which were taken up by the enzyme on rereduction of the metal centers. The proton transfer elicited by ferricyanide represents the net result of deprotonation/protonation reactions linked to anaerobic oxidoreduction of the metal centers. Correction for the ferricyanide-induced pH changes of the proton uptake observed in the oxidation and rereduction phase of the reaction of the reduced oxidase with oxygen gave a measure of the proton consumption in the reduction of O(2) to 2H(2)O. The results show that the expected stoichiometric proton consumption of 4H(+) in the reduction of O(2) to 2H(2)O is differently associated, depending on the actual pH, with the oxidation and reduction phase of COX. Two H(+)/COX are initially taken up in the reduction of O(2) to two OH(-) groups bound to the binuclear Fe a(3)-Cu(B) center. At acidic pHs the third and fourth protons are also taken up in the oxidative phase with formation of 2H(2)O. At alkaline pHs the third and fourth protons are taken up with formation of 2H(2)O only upon rereduction of COX.  相似文献   

9.
Peroxidase from soybean seed coat (SBP) has properties that makes it particularly suited for practical applications. Therefore, it is essential to know its fundamental enzymatic properties. Stopped-flow techniques were used to investigate the pH dependence of the reaction of SBP and hydrogen peroxide. The reaction is linearly dependent on hydrogen peroxide concentration at acidic and neutral pH with the second order rate constant k(1)=2.0x10(7) M(-1) s(-1), pH 4-8. From pH 9.3 to 10.2 the reaction is biphasic, a novel observation for a peroxidase at alkaline pH. A fast reaction has the characteristics of the reaction at neutral pH, and a slow reaction shows hyperbolic dependence on hydrogen peroxide concentration. At pH >10.5 only the slow reaction is seen. The shift in mechanism is coincident with the change in haem iron co-ordination to a six-coordinate low spin hydroxy ligated alkaline form. The pK(a) value for the alkaline transition was observed at 9.7+/-0.1, 9.6+/-0.1 and 9.9+/-0.2 by spectrophotometric titration, the fast phase amplitude, and decrease in the apparent second order rate constant, respectively. An acidic pK(a) at 3.2+/-0.3 was also determined from the apparent second order rate constant. The reactions of soybean peroxidase compounds I and II with veratryl alcohol at pH 2.44 give very similar second order rate constants, k(2)=(2.5+/-0.1)x10(4) M(-1) s(-1) and k(3)=(2.2+/-0.1)x10(4) M(-1) s(-1), respectively, which is unusual. The electronic absorption spectra of compounds I, II and III at pH 7.07 show characteristic bands at 400 and 651 nm (compound I), 416, 527 and 555 nm (compound II), and 414, 541 and 576 nm (compound III). No additional intermediates were observed.  相似文献   

10.
Electrochemical methods were used to activate MnIII and FeIII complexes of meso-tetrakis(N-methyl-4-pyridiniumyl)porphine (H2TMPyP) to cause cleavage of pBR322 DNA and to study their interaction with sonicated calf thymus DNA. Electrochemical reduction of MnIIITMPyP and FeIIITMPyP (at low concentrations) in the presence of O2 was required to activate these complexes. However, FeIIITMPyP at 1 x 10(-6) M produced DNA strand breakage without being electrochemically reduced. At low concentrations, FeIITMPyP was more efficient at cleaving DNA than MnIITMPyP. Reduction of O2 at a platinum electrode also produced some cleavage but to a much smaller extent. The oxidized form of MnIIITMPyP (charge 5+) has higher affinity for sonicated calf thymus (CT) DNA than the reduced form (charge 4+), as determined by the negative shift in E degrees' for the voltammetric wave in the presence of DNA. Both forms of FeIIITMPyP (charge 4+) interact with DNA to about the same extent. Differential pulse voltammetry was used to determine binding constants (K) and binding-site sizes (s) of the interaction of these metalloporphyrins with sonicated CT DNA. The data were analyzed assuming both mobile and static equilibria. MnIIITMPyP binds to DNA (5 mM Tris, 50 mM NaCl, pH 7) with K = 5 (+/- 2) x 10(6) M-1, s = 3 bp (mobile) or K = 3.6 (+/- 0.3) x 10(6) M-1, s = 4 bp (static). FeIIITMPyP at that ionic strength caused DNA precipitation. At higher ionic strength (0.1 M Tris, 0.1 M NaCl, pH 7), FeIIITMPyP associates to DNA with K = 4.4 (+/- 0.2) x 10(4) M-1, s = 5 bp (mobile) or K = 1.9 (+/- 0.1) x 10(4) M-1, s = 6 bp (static).  相似文献   

11.
In the present study, we established a simple and physiologically acceptable in vitro assay system to measure H2O2 generated by human immunoglobulin G (IgG) and other proteins. In addition, the effects of various drugs were also tested in this method. We found that UV irradiation (280 nm) of the test solutions for 1 h at 37 degrees C produced suitable conditions to test the effects of these drugs. The test solution contained 100 microg/ml IgG in 50 mM phosphate buffer (pH 7.4), and 1% dimethylformamide (DMF), a solvent used to dissolve each drug. Phosphate anions were preferable for H2O2 generation. H2O2 concentration in the irradiated sample was determined by continuous photometric measurement of absorption (O.D.) at 340 nm for 600 sec. The decrease in O.D. was due to the oxidation of NADPH by H2O2 mediated by the glutathione redox cycle. H2O2 generation was expressed as O.D.(340 nm decrease/400 sec). IgG (100 microg/ml) generated 6-7 microM H2O2/h. With irradiation, most cytokines, proteins and enzymes failed to generate significant amounts of H2O2. The formation of H2O2 from H2O and UV light-induced singlet oxygen (1O2) was demonstrated by the inhibitory effects of 1O2 quenchers. Dexamethasone (IC50: 6 ng/ml = 1.4x10(-8) M) blocked H2O2 generation catalyzed by IgG. This action was not mediated by binding to the glucocorticoid receptor. Piroxicam (IC50: 20 ng/ml = 6.0 x 10(-6) M) and diclofenac.Na (IC50: 500 ng/ml = 1.6 x 10(-5) M), but not indomethacin, also blocked H2O2 generation. The mechanism underlying the inhibition of IgG-catalyzed H2O2 generation is not clear; however, the possibility exists that these drugs intercept, or interfere with, the approach of water molecules at the catalytic interface(s) of the IgG.  相似文献   

12.
Crystals of K[B(AnErytH(-2)2] x 2 H2O (AnEryt = 1,4-anhydroerythritol) form from aqueous alkaline solutions containing a double molar amount of diol over borate. The spiro-type monoanions are the main borate species in the mother liquors of crystallisation according to 11B and 13C NMR spectroscopy. Ribofuranosides form analogous borate esters using their 1,4-anhydroerythritol core. Crystals of Na[B(Me beta-D-Ribf 2,3H(-2))2] x 2 H2O were grown from alkaline methyl beta-D-ribofuranoside solutions that had attacked boron-containing Duran vessels. NMR spectra show closely resembling borate-ester speciation in solutions of diols with the 1,4-anhydroerythritol core--1,4-anhydroerythritol itself, methyl beta-D-ribofuranoside and guanosine.  相似文献   

13.
The reduction of horse heart ferricytochrome c by the tryptic fragment of bovine liver cytochrome b5 and its dimethyl ester heme (DME)-substituted derivative has been studied as a function of ionic strength, pH, and temperature under solution conditions where the reaction is bimolecular. The rate constant for ferricytochrome c reduction by native ferrocytochrome b5 is 1.8 (+/- 0.2) x 10(7) M-1 s-1 (25 degrees C) with delta H++ = 7.5 (+/- 0.2) kcal/mol and delta S++ = -0.3 (+/- 0.6) eu (pH 7.0, I = 0.348 M). Under the same solution conditions, the reduction of ferricytochrome c by DME-ferrocytochrome b5 proceeds with a rate constant of 1.7 (+/- 0.1) x 10(7) M-1 s-1 with delta H++ = 7.9 (+/- 0.4) kcal/mol and delta S++ = 1 (+/- 1) eu. The rate constants for both reactions are strongly dependent on ionic strength. A detailed electrostatic analysis of the proteins has been performed. Two relatively simple Brownian dynamics simulation models predict rate constants for the reaction between the two native proteins that demonstrate a dependence on ionic strength similar to that observed experimentally. In one of these models, the proteins are treated as spheres with reactive surface patches that are defined by a 5 degrees cone generated about the dipole vector calculated for each protein and aligned with the presumed electron-transfer site near the partially exposed heme edge. The second model replaces the reactive patch assumption with an exponential distance dependence for the probability of reaction that permits estimation of a value for the distance-dependence factor alpha. Calculations with this latter model in combination with the aligned dipole assumption provide a reasonable approximation to the observed ionic strength dependence for the reaction and are consistent with a value of alpha = 1.2 A-1.  相似文献   

14.
Biological effects of ultraviolet A (UVA) irradiation have been ascribed to the photochemical generation of singlet oxygen. Not all effects described in the literature, however, are explicable solely by the generation of singlet oxygen, but rather resemble effects elicited by hydrogen peroxide (H 2 O 2 ). Here, we show that when cells are kept in cell culture media during exposure to UVA, stress kinases, including ERK 1 and ERK 2 as well as Akt (protein kinase B), are activated, whereas there is no or only minor activation when cells are kept in phosphate-buffered saline during irradiation. Indeed, the exposure of cell culture media to UVA (30 J/cm 2 ) results in the generation of significant amounts of H 2 O 2 , with concentrations of about 100 &#119 M. H 2 O 2 concentrations are at least three-fold higher in HEPES-buffered culture media after UVA irradiation. From experiments with solutions of riboflavin, tryptophan or HEPES, as well as combinations thereof, it is concluded that riboflavin mediates the photooxidation of either tryptophan or HEPES, resulting in the generation of H 2 O 2 . Thus, if signaling effects of UVA radiation are to be investigated in cell culture systems, riboflavin and HEPES/tryptophan should be avoided during irradiation because of artificial H 2 O 2 generation. It should be taken into account, however, that in vivo tryptophan and riboflavin might play an important role in the generation of reactive oxygen species by UVA as both substances are abundant in living tissues.  相似文献   

15.
Biological effects of ultraviolet A (UVA) irradiation have been ascribed to the photochemical generation of singlet oxygen. Not all effects described in the literature, however, are explicable solely by the generation of singlet oxygen, but rather resemble effects elicited by hydrogen peroxide (H 2 O 2 ). Here, we show that when cells are kept in cell culture media during exposure to UVA, stress kinases, including ERK 1 and ERK 2 as well as Akt (protein kinase B), are activated, whereas there is no or only minor activation when cells are kept in phosphate-buffered saline during irradiation. Indeed, the exposure of cell culture media to UVA (30 J/cm 2 ) results in the generation of significant amounts of H 2 O 2 , with concentrations of about 100 μM. H 2 O 2 concentrations are at least three-fold higher in HEPES-buffered culture media after UVA irradiation. From experiments with solutions of riboflavin, tryptophan or HEPES, as well as combinations thereof, it is concluded that riboflavin mediates the photooxidation of either tryptophan or HEPES, resulting in the generation of H 2 O 2 . Thus, if signaling effects of UVA radiation are to be investigated in cell culture systems, riboflavin and HEPES/tryptophan should be avoided during irradiation because of artificial H 2 O 2 generation. It should be taken into account, however, that in vivo tryptophan and riboflavin might play an important role in the generation of reactive oxygen species by UVA as both substances are abundant in living tissues.  相似文献   

16.
The apparent cytoplasmic proton diffusion coefficient was measured using pH electrodes and samples of cytoplasm extracted from the giant neuron of a marine invertebrate. By suddenly changing the pH at one surface of the sample and recording the relaxation of pH within the sample, an apparent diffusion coefficient of 1.4 +/- 0.5 x 10(-6) cm2/s (N = 7) was measured in the acidic or neutral range of pH (6.0-7.2). This value is approximately 5x lower than the diffusion coefficient of the mobile pH buffers (approximately 8 x 10(-6) cm2/s) and approximately 68x lower than the diffusion coefficient of the hydronium ion (93 x 10(-6) cm2/s). A mobile pH buffer (approximately 15% of the buffering power) and an immobile buffer (approximately 85% of the buffering power) could quantitatively account for the results at acidic or neutral pH. At alkaline pH (8.2-8.6), the apparent proton diffusion coefficient increased to 4.1 +/- 0.8 x 10(-6) cm2/s (N = 7). This larger diffusion coefficient at alkaline pH could be explained quantitatively by the enhanced buffering power of the mobile amino acids. Under the conditions of these experiments, it is unlikely that hydroxide movement influences the apparent hydrogen ion diffusion coefficient.  相似文献   

17.
Tetrahydrobiopterin ((6R)-5,6,7,8-tetrahydro-L-biopterin (H4biopterin)) is an essential cofactor of nitric-oxide synthases (NOSs), but its role in enzyme function is not known. Binding of the pterin affects the electronic structure of the prosthetic heme group in the oxygenase domain and results in a pronounced stabilization of the active homodimeric structure of the protein. However, these allosteric effects are also produced by the potent pterin antagonist of NOS, 4-amino-H4biopterin, suggesting that the natural cofactor has an additional, as yet unknown catalytic function. Here we show that the 5-methyl analog of H4biopterin, which does not react with O2, is a functionally active pterin cofactor of neuronal NOS. Activation of the H4biopterin-free enzyme occurred in a biphasic manner with half-maximally effective concentrations of approximately 0.2 microM and 10 mM 5-methyl-H4biopterin. Thus, the affinity of the 5-methyl compound was 3 orders of magnitude lower than that of the natural cofactor, allowing the direct demonstration of the functional anticooperativity of the two pterin binding sites of dimeric NOS. In contrast to H4biopterin, which inactivates nitric oxide (NO) through nonenzymatic superoxide formation, up to 1 mM of the 5-methyl derivative did not consume O2 and had no effect on NO steady-state concentrations measured electrochemically with a Clark-type NO electrode. Therefore, reconstitution with 5-methyl-H4biopterin allowed, for the first time, the detection of enzymatic NO formation in the absence of superoxide or NO scavengers. These results unequivocally identify free NO as a NOS product and indicate that reductive O2 activation by the pterin cofactor is not essential to NO biosynthesis.  相似文献   

18.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

19.
The protective role of reactive oxygen scavengers against photodamage was studied in isolated photosystem (PS) I submembrane fractions illuminated (2000 microE x m(-2) x s(-1)) for various periods at 4 degrees C. The photochemical activity of the submembrane fractions measured as P700 photooxidation was significantly protected in the presence of histidine or n-propyl gallate. Chlorophyll photobleaching resulting in a decrease of absorbance and fluorescence, and a blue-shift of both absorbance and fluorescence maximum in the red region, was also greatly delayed in the presence of these scavengers. Western blot analysis revealed the light harvesting antenna complexes of PSI, Lhca2 and Lhca1, were more susceptible to strong light when compared to Lhca3 and Lhca4. The reaction-center proteins PsaB, PsaC, and PsaE were most sensitive to strong illumination while other polypeptides were less affected. Addition of histidine or n-propyl gallate lead to significant protection of reaction-center proteins as well as Lhca against strong illumination. Circular dichroism (CD) spectra revealed that the alpha-helix content decreased with increasing period of light exposure, whereas beta-strands, turns, and unordered structure increased. This unfolding was prevented with the addition of histidine or n-propyl gallate even after 10 h of strong illumination. Catalase or superoxide dismutase could not minimize the alteration of PSI photochemical activity and structure due to photodamage. The specific action of histidine and n-propyl gallate indicates that 1O2 was the main form of reactive oxygen species responsible for strong light-induced damage in PSI submembrane fractions.  相似文献   

20.
The functional properties of the recombinant C-terminal dimerization domain of the Pseudomonas aeruginosa Fur (ferric uptake regulator) protein expressed in and purified from Escherichia coli have been evaluated. Sedimentation velocity measurements demonstrate that this domain is dimeric, and the UV CD spectrum is consistent with a secondary structure similar to that observed for the corresponding region of the crystallographically characterized wild-type protein. The thermal stability of the domain as determined by CD spectroscopy decreases significantly as pH is increased and increases significantly as metal ions are added. Potentiometric titrations (pH 6.5) establish that the domain possesses a high-affinity and a low-affinity binding site for metal ions. The high-affinity (sensory) binding site demonstrates association constants (K(A)) of 10(+/-7)x10(6), 5.7(+/-3)x10(6), 2.0(+/-2)x10(6) and 2.0(+/-3)x10(4) M(-1) for Ni2+, Zn2+, Co2+ and Mn2+ respectively, while the low-affinity (structural) site exhibits association constants of 1.3(+/-2)x10(6), 3.2(+/-2)x10(4), 1.76(+/-1)x10(5) and 1.5(+/-2)x10(3) M(-1) respectively for the same metal ions (pH 6.5, 300 mM NaCl, 25 degrees C). The stability of metal ion binding to the sensory site follows the Irving-Williams order, while metal ion binding to the partial sensory site present in the domain does not. Fluorescence experiments indicate that the quenching resulting from binding of Co2+ is reversed by subsequent titration with Zn2+. We conclude that the domain is a reasonable model for many properties of the full-length protein and is amenable to some analyses that the limited solubility of the full-length protein prevents.  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