首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Jensen FB 《The FEBS journal》2008,275(13):3375-3387
The nitrite reductase activity of deoxyhemoglobin has received much recent interest because the nitric oxide produced in this reaction may participate in blood flow regulation during hypoxia. The present study used spectral deconvolution to characterize the reaction of nitrite with carp and rabbit hemoglobin at different constant oxygen tensions that generate the full range of physiological relevant oxygen saturations. Carp is a hypoxia-tolerant species with very high hemoglobin oxygen affinity, and the high R-state character and low redox potential of the hemoglobin is hypothesized to promote NO generation from nitrite. The reaction of nitrite with deoxyhemoglobin leads to a 1 : 1 formation of nitrosylhemoglobin and methemoglobin in both species. At intermediate oxygen saturations, the reaction with deoxyhemoglobin is clearly favored over that with oxyhemoglobin, and the oxyhemoglobin reaction and its autocatalysis are inhibited by nitrosylhemoglobin from the deoxyhemoglobin reaction. The production of NO and nitrosylhemoglobin is faster and higher in carp hemoglobin with high O(2) affinity than in rabbit hemoglobin with lower O(2) affinity, and it correlates inversely with oxygen saturation. In carp, NO formation remains substantial even at high oxygen saturations. When oxygen affinity is decreased by T-state stabilization of carp hemoglobin with ATP, the reaction rates decrease and NO production is lowered, but the deoxyhemoglobin reaction continues to dominate. The data show that the reaction of nitrite with hemoglobin is dynamically influenced by oxygen affinity and the allosteric equilibrium between the T and R states, and that a high O(2) affinity increases the nitrite reductase capability of hemoglobin.  相似文献   

2.
Recent studies have detected a (1)H nuclear magnetic resonance (NMR) reporter signal of metmyoglobin (metMb) during bradykinin stimulation of an isolated mouse heart. The observation has led to the hypothesis that Mb reacts with cellular nitric oxide (NO). However, the hypothesis depends on an unequivocal detection of metMb signals in vivo. In solution, nitrite oxidization of Mb produces a characteristic set of paramagnetically shifted (1)H NMR signals. In the upfield spectral region, MbO(2) and MbCO exhibit the gammaCH(3) Val E11 signals at -2.8 and -2.4 ppm, respectively. In the same spectral region, nitrite oxidation of Mb produces a set of signals at -3.7 and -4.7 ppm at 35 degrees C. Previous studies have confirmed the visibility of metMb signals in perfused rat myocardium. With bradykinin infusion, perfusion pressure and rate-pressure product decrease, consistent with endogenous NO formation. However, neither myocardial O(2) consumption nor high-energy phosphate levels, as reflected in the (31)P NMR signals, show any significant change. Bradykinin still triggers a similar physiological response even in the presence of CO that is sufficient to inhibit 86% Mb. In all cases, the (1)H NMR spectra from perfused rat myocardium reveal no metMb signals. The results suggest that bradykinin-induced NO does not interact significantly with cellular Mb to produce an NMR-detectable quantity of metMb in the perfused rat myocardium. As a consequence, the experiments cannot confirm the intriguing proposal that Mb acts as a cellular NO scavenger.  相似文献   

3.
Myoglobin (Mb) serves in the facilitated diffusion and storage of O? in heart and skeletal muscle, where it also regulates O? consumption via nitric oxide (NO) scavenging or generation. S-nitrosation at reactive cysteines may generate S-nitroso Mb (Mb-SNO) and contribute further to NO homeostasis. In being a monomer, Mb is commonly believed to lack allosteric control of heme reactivity. Here, we test whether in rainbow trout, a fast swimmer living in well-aerated water, the Mb-O? affinity is regulated by ionic cofactors and S-nitrosation. O? equilibria showed the lowest O? affinity ever reported among vertebrate Mbs (P?? = 4.92 ± 0.29 mmHg, 25°C), a small overall heat of oxygenation (ΔH = -12.03 kcal/mol O?), and no effect of chloride, pH, or lactate. Although the reaction with 4,4'-dithiodipyridine (4-PDS) showed 1.3-1.9 accessible thiols per heme, the reaction of Mb with S-nitroso cysteine (Cys-NO) and S-nitrosoglutathione (GSNO) to generate Mb-SNO yielded ~0.3-0.6 and ~0.1 SNO/heme, respectively, suggesting S-nitrosation at only one cysteine (likely Cys1?). At ~60% S-nitrosation, trout Mb-SNO showed a higher O? affinity (P?? = 2.23 ± 0.19 mmHg, 20°C) than unmodified Mb (3.36 ± 0.11 mmHg, 20°C). Total SNO levels measured by chemiluminescence in trout myocardial preparations decreased after hypoxia, but not significantly, indicating that transnitrosation reactions between thiols may occur in vivo. Our data reveal a novel, S-nitrosation-dependent allosteric mechanism in this low-affinity Mb that may contribute to targeted O?-linked SNO release in the hypoxic fish heart and be of importance in preserving cardiac function during intense exercise.  相似文献   

4.
Site-specific mutants of human myoglobin (Mb) have been prepared, in which Leu29 (B10) is replaced by Ala(L29A) or Ile(L29I), in order to examine the influence of this highly conserved residue in the hydrophobic clusters of the heme distal site on the heme environmental structure and ligand binding properties of Mb. Structural characterizations of these recombinant Mbs are studied by electronic absorption, infrared (IR), one- and two-dimensional proton nuclear magnetic resonance spectroscopies, and ligand-binding kinetics by laser photolysis measurements under ambient and high pressures (up to 2000 bar). Multiple split carbon monoxide (CO) stretch bands in the IR spectra of mutant Mbs exhibit a relative decrease of the 1945 cm-1 band (approximately 50%) which is associated with an upright binding geometry of CO, accompanied by an increase of the tilted CO conformer at 1932 cm-1. On the basis of these results, replacement of Leu29(B10) by Ala or Ile appears to allow bound CO to rotate from a conformation pointing toward the beta meso carbon of the heme group to the one pointing toward the alpha meso carbon atom, presumably filling the space left by removal of the delta 2 carbon atom of Leu29(B10). These substitutions cause the rate constants for CO and O2 association to decrease almost 3-5-fold. Present results show that CO and O2 bindings to the heme iron of Mb are controlled by Leu29(B10) by influencing the structure of close vicinity of the heme and the geometry of iron-bound ligand. Further, mutant Mbs (Leu72(E15)----Ala and Leu104 (G5)----Ala) which have altered residues in another hydrophobic clusters around proximal and distal site are also examined.  相似文献   

5.
Cao W  Christian JF  Champion PM  Rosca F  Sage JT 《Biochemistry》2001,40(19):5728-5737
Flash photolysis investigations of horse heart metmyoglobin bound with NO (Mb(3+)NO) reveal the kinetics of water entry and binding to the heme iron. Photodissociation of NO leaves the sample in the dehydrated Mb(3+) (5-coordinate) state. After NO photolysis and escape, a water molecule enters the heme pocket and binds to the heme iron, forming the 6-coordinate aquometMb state (Mb(3+)H2O). At longer times, NO displaces the H2O ligand to reestablish equilibrium. At 293 K, we determine a value k(w) approximately 5.7 x 10(6) s(-1) for the rate of H2O binding and estimate the H2O dissociation constant as 60 mM. The Arrhenius barrier height H(w) = 42 +/- 3 kJ/mol determined for H2O binding is identical to the barrier for CO escape after photolysis of Mb(2+)CO, within experimental uncertainty, consistent with a common mechanism for entry and exit of small molecules from the heme pocket. We propose that both processes are gated by displacement of His-64 from the heme pocket. We also observe that the bimolecular NO rebinding rate is enhanced by 3 orders of magnitude both for the H64L mutant, which does not bind water, and for the H64G mutant, where the bound water is no longer stabilized by hydrogen bonding with His-64. These results emphasize the importance of the hydrogen bond in stabilizing H2O binding and thus preventing NO scavenging by ferric heme proteins at physiological NO concentrations.  相似文献   

6.
1H-NMR spectra of deoxy myoglobins (Mbs) from shark (Galeorhinus japonicus), horse, and sperm whale have been studied to gain insights into their active site structure. It has been demonstrated for the first time that nuclear Overhauser effect (NOE) can be observed between heme peripheral side-chain proton resonances of these paramagnetic complexes. Val-E11 methyl and His-F8 C delta H proton resonances of these Mbs were also assigned from the characteristic shift and line width. The hyperfine shift of the former resonance was used to calculate the magnetic anisotropy of the protein. The shift analysis of the latter resonance, together with the previously assigned His-F8 N delta H proton resonance, revealed that the strain on the Fe-N epsilon bond is in the order horse Mb approximately whale Mb < shark Mb and that the hydrogen bond strength of the His-F8 N delta H proton to the main-chain carbonyl oxygen in the preceding turn of the F helix is in the order shark Mb < horse Mb < whale Mb. Weaker Feporphyrin interaction in shark Mb was manifested in a smaller shift of the heme methyl proton resonance and appears to result from distortion of the coordination geometry in this Mb. Larger strain on the Fe-N epsilon bond in shark Mb should be to some extent attributed to its lowered O2 affinity (P50 = 1.1 mmHg at 20 degrees C), compared to whale and horse Mbs.  相似文献   

7.
Hirota S  Azuma K  Fukuba M  Kuroiwa S  Funasaki N 《Biochemistry》2005,44(30):10322-10327
Human myoglobin (Mb) possesses a unique cysteine (Cys110), whereas other mammalian Mbs do not. To investigate the effect of a cysteine residue on Mb, we introduced cysteine to various sites on the surface of sperm whale Mb (K56C, V66C, K96C, K102C, A125C, and A144C) by mutation. The cysteines were inserted near the end of alpha-helices, except for V66C, where the cysteine was introduced in the middle of an alpha-helix. Reduction of the heme was observed for each mutant metMb by incubation at 37 degrees C under carbon monoxide atmosphere, which was much faster than reduction of wild-type metMb under the same condition. Heme reduction did not occur significantly under nitrogen or oxygen atmospheres. The rate constant for heme reduction increased for higher mutant Mb concentration, whereas it did not change significantly when the CO concentration was reduced from 100% CO to 50% CO with 50% O(2). The similarity in the rate constants with different CO concentrations indicates that CO stabilizes the reduced heme by coordination to the heme iron. SDS-PAGE analysis showed that mutant Mb dimers were formed by incubation under CO atmosphere but not under air. These dimers were converted back to Mb monomers by an addition of 2-mercaptoethanol, which showed formation of a Mb dimer through a disulfide bond. The rate constant decreased in general as the heme-cysteine distance was increased, although V66C Mb exhibited a very small rate constant. Since V66 is placed in the middle of an alpha-helix, steric hindrance would occur and prevent formation of a dimer when the cysteine residues of two different V66C Mb molecules interact with each other. The rate constants also decreased for K56C and A144C Mbs presumably because of the electrostatic repulsion during dimer formation, since they are relatively charged around the inserted cysteine.  相似文献   

8.
Nitrite is now recognized as a storage pool of bioactive nitric oxide (NO). Hemoglobin (Hb) and myoglobin (Mb) convert, under certain conditions, nitrite to NO. This newly discovered nitrite reductase activity of Hb and Mb provides an attractive alternative to mammalian NO synthesis from the NO synthase pathway that requires dioxygen. We recently reported the X-ray crystal structure of the nitrite adduct of ferric horse heart Mb, and showed that the nitrite ligand binds in an unprecedented O-binding (nitrito) mode to the d(5) ferric center in Mb(III)(ONO) [D.M. Copeland, A. Soares, A.H. West, G.B. Richter-Addo, J. Inorg. Biochem. 100 (2006) 1413-1425]. We also showed that the distal pocket in Mb allows for different conformations of the NO ligand (120 degrees and 144 degrees ) in Mb(II)NO depending on the mode of preparation of the compound. In this article, we report the crystal structures of the nitrite and NO adducts of manganese-substituted hh Mb (a d(4) system) and of the nitrite adduct of cobalt-substituted hh Mb (a d(6) system). We show that the distal His64 residue directs the nitrite ligand towards the rare nitrito O-binding mode in Mn(III)Mb and Co(III)Mb. We also report that the distal pocket residues allow a stabilization of an unprecendented bent MnNO moiety in Mn(II)MbNO. These crystal structural data, when combined with the data for the aquo, methanol, and azide MnMb derivatives, provide information on the role of distal pocket residues in the observed binding modes of nitrite and NO ligands to wild-type and metal-substituted Mb.  相似文献   

9.
Five Tn5 mutants of Pseudomonas fluorescens AK-15 deficient in dissimilatory reduction of nitrite were isolated and characterized. Two insertions occurred inside the nitrite reductase structural gene (nirS) and resulted in no detectable nitrite reductase protein on a Western immunoblot. One mutant had Tn5 inserted inside nirC, the third gene in the same operon, and produced a defective nitrite reductase protein. Two other mutants had insertions outside of this nir operon and also produced defective proteins. All of the Nir- mutants characterized showed not only loss of nitrite reductase activity but also a significant decrease in nitric oxide reductase activity. When cells were incubated with 15NO in H2(18)O, about 25% of the oxygen found in nitrous oxide exchanged with H2O. The extent of exchange remained constant throughout the reaction, indicating the incorporation of 18O from H2(18)O reached equilibrium rapidly. In all nitrite reduction-deficient mutants, less than 4% of the 18O exchange was found, suggesting that the hydration and dehydration step was altered. These results indicate that the factors involved in dissimilatory reduction of nitrite influenced the subsequent NO reduction in this organism.  相似文献   

10.
The dissimilatory nitrite reductase (cytochrome c,d1) from Pseudomonas aeruginosa was observed at pH 7.5 to catalyze nitrosyl transfer (nitrosation) between [15N]nitrite and several N-nucleophiles or H2 18O, with rate enhancement of the order of 10(8) relative to analogous chemical reactions. The reducing system (ascorbate, N,N,N',N'-tetramethylphenylenediamine) could reduce nitrite (but not NO) enzymatically and had essentially no direct chemical reactivity toward nitrite or NO. The N-nitrosations showed saturation kinetics with respect to the nucleophile and, while exhibiting Vmax values which varied by about 40-fold, nevertheless showed little or no dependence of Vmax on nucleophile pKa. The N-nitrosations and NO-2/H2O-18O exchange required the reducing system, whereas NO/H2O-18O exchange was inhibited by the reducing system. NO was not detected to serve as a nitrosyl donor to N-nucleophiles. These and other kinetic observations suggest that the enzymatic nitrosyl donor is an enzyme-bound species derived from reduced enzyme and one molecule of nitrite, possibly a heme-nitrosyl compound (E-FeII X NO+) for which there is precedence. Nitrosyl transfer to N-nucleophiles may occur within a ternary complex of enzyme, nitrite, and nucleophile. Catalysis of nitrosyl transfer by nitrite reductase represents a new class of enzymatic reactions and may present another example of electrophilic catalysis by a metal center. The nitrosyl donor trapped by these reactions is believed to represent an intermediate in the reduction of nitrite by cytochrome c,d1.  相似文献   

11.
Nitrite reacts with deoxyhemoglobin to generate nitric oxide (NO). This reaction has been proposed to contribute to nitrite-dependent vasodilation in vivo and potentially regulate physiological hypoxic vasodilation. Paradoxically, while deoxyhemoglobin can generate NO via nitrite reduction, both oxyhemoglobin and deoxyhemoglobin potently scavenge NO. Furthermore, at the very low O(2) tensions required to deoxygenate cell-free hemoglobin solutions in aortic ring bioassays, surprisingly low doses of nitrite can be reduced to NO directly by the blood vessel, independent of the presence of hemoglobin; this makes assessments of the role of hemoglobin in the bioactivation of nitrite difficult to characterize in these systems. Therefore, to study the O(2) dependence and ability of deoxhemoglobin to generate vasodilatory NO from nitrite, we performed full factorial experiments of oxyhemoglobin, deoxyhemoglobin, and nitrite and found a highly significant interaction between hemoglobin deoxygenation and nitrite-dependent vasodilation (P < or = 0.0002). Furthermore, we compared the effect of hemoglobin oxygenation on authentic NO-dependent vasodilation using a NONOate NO donor and found that there was no such interaction, i.e., both oxyhemoglobin and deoxyhemoglobin inhibited NO-mediated vasodilation. Finally, we showed that another NO scavenger, 2-carboxyphenyl-4,4-5,5-tetramethylimidazoline-1-oxyl-3-oxide, inhibits nitrite-dependent vasodilation under normoxia and hypoxia, illustrating the uniqueness of the interaction of nitrite with deoxyhemoglobin. While both oxyhemoglobin and deoxyhemoglobin potently inhibit NO, deoxyhemoglobin exhibits unique functional duality as an NO scavenger and nitrite-dependent NO generator, suggesting a model in which intravascular NO homeostasis is regulated by a balance between NO scavenging and NO generation that is dynamically regulated by hemoglobin's O(2) fractional saturation and allosteric nitrite reductase activity.  相似文献   

12.
Dai Z  Xu X  Ju H 《Analytical biochemistry》2004,332(1):23-31
The direct electrochemistry of myoglobin (Mb) immobilized on a hexagonal mesoporous silica (HMS)-modified glassy carbon electrode was described. The interaction between Mb and HMS was investigated by using Fourier transfer infrared spectroscopy, nitrogen adsorption isotherm, and cyclic voltammetry. Two couples of redox peaks corresponding to Fe(III) to Fe(II) conversion of the Mb intercalated in the mesopores and adsorbed on the surface of the HMS were observed with the formal potentials of -0.167 and -0.029V in 0.1M, pH 7.0, phosphate buffer solution, respectively. The electrode reaction showed a surface-controlled process with one proton transfer. The immobilized Mb displayed good electrocatalytic responses to the reduction of both hydrogen peroxide (H(2)O(2)) and nitrite (NO(2)(-)), which were used to develop novel sensors for H(2)O(2) and NO(2)(-). The apparent Michaelis-Menten constants of the immobilized Mb for H(2)O(2) and NO(2)(-) were 0.065 and 0.72mM, respectively, showing good affinity. Under optimal conditions, the sensors could be used for the determinations of H(2)O(2) ranging from 4.0 to 124microM and NO(2)(-) ranging from 8.0 to 216microM. The detection limits were 6.2x10(-8) and 8.0x10(-7)M at 3 sigma, respectively. The HMS provided a novel matrix for protein immobilization and the construction of biosensors via the direct electron transfer of immobilized protein.  相似文献   

13.
1. Cells of Nitrosomonas europaea produced N(2)O during the oxidation of ammonia and hydroxylamine. 2. The end-product of ammonia oxidation, nitrite, was the predominant source of N(2)O in cells. 3. Cells also produced N(2)O, but not N(2) gas, by the reduction of nitrite under anaerobic conditions. 4. Hydroxylamine was oxidized by cell-free extracts to yield nitrite and N(2)O aerobically, but to yield N(2)O and NO anaerobically. 5. Cell extracts reduced nitrite both aerobically and anaerobically to NO and N(2)O with hydroxylamine as an electron donor. 6. The relative amounts of NO and N(2)O produced during hydroxylamine oxidation and/or nitrite reduction are dependent on the type of artificial electron acceptor utilized. 7. Partially purified hydroxylamine oxidase retained nitrite reductase activity but cytochrome oxidase was absent. 8. There is a close association of hydroxylamine oxidase and nitrite reductase activities in purified preparations.  相似文献   

14.
Systemic hypoxia results in oxidative stress due to a change in the reactive oxygen species (ROS)-nitric oxide (NO) balance. These experiments explored two mechanisms for the altered ROS-NO balance: 1) decreased NO synthesis by NO synthase due to limited O(2) substrate availability and 2) increased superoxide generation. ROS levels and leukocyte adherence in mesenteric venules of rats during hypoxia were studied in the absence and presence of an NO donor [spermine NONOate (SNO)] and of the NO precursor L-arginine. We hypothesized that if the lower NO levels during hypoxia were due to O(2) substrate limitation, L-arginine would not prevent hypoxia-induced microvascular responses. Graded hypoxia (produced by breathing 15, 10, and 7.5% O(2)) increased both ROS (123 +/- 6, 148 +/- 11, and 167 +/- 3% of control) and leukocyte adherence. ROS levels during breathing of 10 and 7.5% O(2) were significantly attenuated by SNO (105 +/- 6 and 108 +/- 3%, respectively) and L-arginine (117 +/- 5 and 115 +/- 2%, respectively). Both interventions reduced leukocyte adherence by similar degrees. The fact that the effects of L-arginine were similar to those of SNO does not support the idea that NO generation is impaired in hypoxia and suggests that tissue NO levels are depleted by the increased ROS during hypoxia.  相似文献   

15.
Bacterial denitrification reverses nitrogen fixation in the global N-cycle by transforming nitrate or nitrite to dinitrogen. Both nitrite and nitric oxide (NO) are considered as the chemical species within the denitrification pathway, that precede nitrous oxide (N2O), the first recognized intermediate with N,N-bonds antecedent to N2. Molecular cloning of the structural genes for NO reductase from Pseudomonas stutzeri has allowed us to generate the first mutants defective in NO utilization (Nor- phenotype) by marker exchange of the norCB genes with a gene cassette for gentamicin resistance. Nitric oxide reductase was found to be an indispensable component for denitrification; its loss constituted a conditionally lethal mutation. NO as the sole product accumulated from nitrite by mutant cells induced for nitrite respiration (denitrification). The Nor- mutant lost the capability to reduce NO and did not grow anymore anaerobically on nitrate. A Nir-Nor- double mutation, that inactivated also the respiratory nitrite reductase cytochrome cd1 rendered the bacterium again viable under anaerobiosis. Our observations provide evidence for a denitrification pathway in vivo of NO2(-)----NO----N2O, and N,N-bond formation catalyzed by NO reductase and not by cytochrome cd1.  相似文献   

16.
Both carbon monoxide (CO) and nitric oxide (NO) play fundamental roles in plant responses to environmental stress. Glutathione (GSH) homeostasis through the glutathione-ascorbate cycle regulates the cellular redox status and protects the plant from damage due to reactive oxygen species (ROS) or reactive nitrogen species (RNS). Most recalcitrant seeds are sensitive to chilling stress, but the roles of and cross talk among CO, NO, ROS, and GSH in recalcitrant seeds under low temperature are not well understood. Here, we report that the germination of recalcitrant Baccaurea ramiflora seeds shows sensitivity to chilling stress, but application of exogenous CO or NO markedly increased GSH accumulation, enhanced the activities of antioxidant enzymes involved in the glutathione-ascorbate cycle, decreased the content of H(2)O(2) and RNS, and improved the tolerance of seeds to low-temperature stress. Compared to orthodox seeds such as maize, only transient accumulation of CO and NO was induced and only a moderate increase in GSH was shown in the recalcitrant B. ramiflora seeds. Exogenous CO or NO treatment further increased the GSH accumulation and S-nitrosoglutathione reductase (GSNOR) activity in B. ramiflora seeds under chilling stress. In contrast, suppressing CO or NO generation, removing GSH, or blocking GSNOR activity resulted in increases in ROS and RNS and impaired the germination of CO- or NO-induced seeds under chilling stress. Based on these results, we propose that CO acts as a novel regulator to improve the tolerance of recalcitrant seeds to low temperatures through NO-mediated glutathione homeostasis.  相似文献   

17.
The mechanism of the denitrification and nitrosation reactions catalyzed by the heme cd-containing nitrite reductase from Pseudomonas stutzeri JM 300 has been studied with whole cell suspensions using H2(18)O, 15NO, and 15NO-2. The extent of H2(18)O exchange with the enzyme-bound nitrosyl intermediate, as determined by the 18O content of product N2O, decreased with increasing nitrite concentration, which is consistent with production of N2O by sequential reaction of two nitrite ions with the enzyme. Reaction of NO with whole cells in H2(18)O gave amounts of 18O in the N2O product consistent with equilibration of nitric oxide with a small pool of free nitrite. Using 15NO and NH2OH, competition between denitrification and nitrosation reactions was demonstrated, as is required if the enzyme-nitrosyl complex is an intermediate in both nitrosation and denitrification reactions. The first evidence for exchange of 18O between H2(18)O and a nitrosation intermediate occurring after the enzyme-nitrosyl complex, presumably an enzyme-bound nitrosamine, has been obtained. The collective results are most consistent with denitrification N2O originating via attack of NO-2 on a coordinated nitrosyl, as proposed earlier (Averill, B. A., and Tiedje, J. M. (1982) FEBS Lett. 138, 8-11).  相似文献   

18.
Brunori M  Gibson QH 《EMBO reports》2001,2(8):674-679
Small globular proteins contain internal cavities and packing defects that reduce thermodynamic stability but seem to play a role in controlling function by defining pathways for the diffusion of the ligand/substrate to the active site. In the case of myoglobin (Mb), a prototype for structure–function relationship studies, the photosensitivity of the adduct of the reduced protein with CO, O2 and NO allows events related to the migration of the ligand through the matrix to be followed. The crystal structures of intermediate states of wild-type (wt) and mutant Mbs show the photolysed CO to be located either in the distal heme pocket (primary docking site) or in one of two alternative cavities (secondary docking sites) corresponding to packing defects accessible to an atom of xenon. These results convey the general picture that pre-existing internal cavities are involved in controlling the dynamics and reactivity of the reactions of Mb with O2 and other ligands, including NO.  相似文献   

19.
Some recent studies on the pathway of nitrogen and the reductases of denitrification are reviewed. The available evidence suggests that while the intermediates of denitrification can remain enzyme-bound (presumably to nitrite reductase) prior to formation of N2O, NO and nitroxyl (HNO) can be released in part by certain bacteria. Release of NO is recognized by a nitrite/NO?15N exchange reaction and isotopic scrambling in product N2O; release of nitroxyl by Pseudomonas stutzeri is recognized by isotopic scrambling of nitrite and NO in product N2O in absence of exchange and affords evidence that the first N?N bond forms in denitrification at the N1+ redox level. The recent purification and partial characterization of nitrous oxide reductase are described. The ability of the dissimilatory nitrite reductase to activate nitrite for nitrosyl transfer affords a new chemical probe into the mechanism of action of this central enzyme. It would appear that reduction of nitrite is subject to electrophilic catalysis. 18O studies show that dissociation of nitrite from nitrite reductase can be slow relative to competing reduction or nitrosyl transfer.  相似文献   

20.
Tn5 was used to generate mutants that were deficient in the dissimilatory reduction of nitrite for Pseudomonas sp. strain G-179, which contains a copper nitrite reductase. Three types of mutants were isolated. The first type showed a lack of growth on nitrate, nitrite, and nitrous oxide. The second type grew on nitrate and nitrous oxide but not on nitrite (Nir-). The two mutants of this type accumulated nitrite, showed no nitrite reductase activity, and had no detectable nitrite reductase protein bands in a Western blot (immunoblot). Tn5 insertions in these two mutants were clustered in the same region and were within the structural gene for nitrite reductase. The third type of mutant grew on nitrate but not on nitrite or nitrous oxide (N2O). The mutant of this type accumulated significant amounts of nitrite, NO, and N2O during anaerobic growth on nitrate and showed a slower growth rate than the wild type. Diethyldithiocarbamic acid, which inhibited nitrite reductase activity in the wild type, did not affect NO reductase activity, indicating that nitrite reductase did not participate in NO reduction. NO reductase activity in Nir- mutants was lower than that in the wild type when the strains were grown on nitrate but was the same as that in the wild type when the strains were grown on nitrous oxide. These results suggest that the reduction of NO and N2O was carried out by two distinct processes and that mutations affecting nitrite reduction resulted in reduced NO reductase activity following anaerobic growth with nitrate.  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