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 146 毫秒
1.
Fisher K  Dilworth MJ  Newton WE 《Biochemistry》2006,45(13):4190-4198
The Mo-nitrogenase-catalyzed reduction of both cyanide and azide results in the production of excess NH3, which is an amount of NH3 over and above that expected to be formed from the well-recognized reactions. Several suggestions about the possible sources of excess NH3 have been made, but previous attempts to characterize these reactions have met with either limited (or no) success or controversy. Because V-nitrogenase has a propensity to release partially reduced intermediates, e.g., N2H4 during N2 reduction, it was selected to probe the reduction of cyanide and azide. Sensitive assay procedures were developed and employed to monitor the production of either HCHO or CH3OH (its further two-electron-reduced product) from HCN. Like Mo-nitrogenase, V-nitrogenase suffered electron-flux inhibition by CN- (but was much less sensitive than Mo-nitrogenase), but unlike the case for Mo-nitrogenase, MgATP hydrolysis was also inhibited by CN-. V-Nitrogenase also released more of the four-electron-reduced intermediate, CH3NH2, than did Mo-nitrogenase. At high NaCN concentrations, V-nitrogenase directed a significant percentage of electron flux into excess NH3, and under these conditions, substantial amounts of HCHO, but no CH3OH, were detected for the first time. With azide, in contrast to the case for Mo-nitrogenase, both total electron flux and MgATP hydrolysis with V-nitrogenase were inhibited. V-Nitrogenase, unlike Mo-nitrogenase, showed no preference between the two-electron reduction to N2-plus-NH3 and the six-electron reduction to N2H4-plus-NH3. V-Nitrogenase formed more excess NH3, but reduction of the N2 produced by the two-electron reduction of N3(-) was not its source. Rather, it was formed directly by the eight-electron reduction of N3(-). Unlike Mo-nitrogenase, CO could not completely eliminate either cyanide or azide reduction by V-nitrogenase. CO did, however, eliminate the inhibition of both electron flux and MgATP hydrolysis by CN-, but not that caused by azide. These different responses to CO suggest different sites or modes of interaction for these two substrates with V-nitrogenase.  相似文献   

2.
Tolland JD  Thorneley RN 《Biochemistry》2005,44(27):9520-9527
Stopped-flow FTIR spectroscopy was used to monitor continuously the pre-steady- and steady-state phases of azide reduction by nitrogenase and the accompanying hydrolysis of ATP. This was characterized by a ca. 1.3 s lag phase that is explained by the number of Fe protein cycles required to effect the reductions of azide to N(2) + NH(3), N(2)H(4) + NH(3), or 3NH(3). Extrapolation of the steady-state time course for azide reduction to zero time showed that one azide binds within 200 ms to each FeMo cofactor. Inhibition of azide reduction by CO was established at times <400 ms, which was faster than the appearance of the first observable IR band assigned to CO (1904 cm(-)(1) detectable at ca. 1 s with maximum amplitude at ca. 7 s). IR bands associated with the rapidly formed (<400 ms) CO species that inhibits azide reduction were not observed over the range 1700-2100 cm(-)(1). This suggests either that the CO is initially bridging two or more Fe atoms or that a rapid reduction of CO to a formyl state occurs by insertion into a metal-hydride bond. The frequencies and time courses for the appearance and loss of the CO bands under hi- and lo-CO conditions were essentially unaffected by the presence of 20 mM azide, consistent with CO being a noncompetitive inhibitor of azide reduction and with azide and CO binding to different sites on the FeMo cofactor.  相似文献   

3.
Nitrogenase reactivity: methyl isocyanide as substrate and inhibitor   总被引:3,自引:0,他引:3  
We have examined the interaction of methyl isocyanide with the purified component proteins of Azotobacter vinelandii nitrogenase (Av1 and Av2). CH3NC was shown to be a potent reversible inhibitor (Ki = 158 microM) of total electron flow, apparently uncoupling magnesium adenosine 5'-triphosphate hydrolysis from electron transfer to substrate. CH3NC is a substrate (Km = 0.688 mM at Av2/Av1 = 8), and extrapolation of the data indicates that at high enough CH3NC concentration, H2 evolution can be eliminated. The products are methane plus methylamine (six electrons) and dimethylamine (four electrons). There is an excess (relative to methane) of methylamine formed, which may arise by hydrolysis of a two-electron intermediate. A rapid high-performance liquid chromatography/fluorescence method was developed for methylamine determination. The products C2H4 and C2H6 appear to be formed via a reduction followed by an insertion mechanism. CH3NC appears to be reduced at an enzyme state more oxidized than the one responsible for H2 evolution or N2 reduction. Other substrates (C2H2 greater than N2 congruent to azide greater than N2O) all both relieve CH3NC inhibition and inhibit CH3NC reduction. Both effects occur in the same relative order, implying productive (substrate) and nonproductive (inhibitor) modes of binding of CH3NC to the same site.  相似文献   

4.
The reduction of N2 to 2NH3 by Klebsiella pneumoniae nitrogenase was studied by a rapid-quench technique. The pre-steady-state time course for N2H4, formed on quenching by the acid-induced hydrolysis of an enzyme-bound intermediate in N2 reduction, showed a 230 ms lag followed by a damped oscillatory approach to a constant concentration in the steady state. The pre-steady-state time course for NH3 formation exhibited a lag of 500 ms and a burst phase that was essentially complete at 1.5s, before a steady-state rate was achieved. These time courses have been simulated by using a previously described kinetic model for the mechanism of nitrogenase action [Lowe & Thorneley (1984) Biochem. J. 224, 877-886]. A hydrazido(2-) structure (=N-NH2) is favoured for the intermediate that yields N2H4 on quenching. The NH3-formation data indicate enzyme-bound metallo-nitrido (identical to N) or -imido (=NH) intermediates formed after N-N bond cleavage to produce the first molecule of NH3 and which subsequently give the second molecule of NH3 by hydrolysis on quenching. The simulations require stoichiometric reduction of one N2 molecule at each Mo and the displacement of one H2 when N2 binds to the MoFe protein. Inhibition by H2 of N2-reduction activity occurs before the formation of the proposed hydrazido(2-) species, and is explained by H2 displacement of N2 at the active site.  相似文献   

5.
Two types of mechanisms for the enzymatic reduction of NO2- to N2O have been proposed. In one, two NO2- ions are reduced in parallel, with the nitrogen-nitrogen bond formed from reduced intermediates. In the second, the two NO2- ions enter the reaction sequentially, with the nitrogen of at least one of the two having a valence of 3+ when the nitrogen-nitrogen bond is formed. Our objective was to distinguish between these two types of mechanism. Toward that end, the exchange of 18O from H2O to NO2- and the overall nitrogen isotopic fractionation factor (beta obs) were measured. The rate of exchange of oxygen from H2O to NO2-, resulting from a protonation-dehydration step preceding reductive events in both mechanisms, was less than 10% of the rate of denitrification at both low and high [NO2-]. The value of beta obs was 1.010 +/- 0.001 and 1.020 +/- 0.001 at low and high [NO2-], respectively. Expressions for beta obs, as a function of the measured rate of entry of oxygen from H2O into NO2-, were derived for both types of mechanism. The measured dependence of beta obs on substrate concentration, as constrained by the 18O exchange data, is inconsistent with the first type of mechanism, but consistent with the second type. Thus, by combining nitrogen isotopic fractionation and 18O exchange data, we rule out any mechanism in Pseudomonas stutzeri in which NO2- ions are reduced in parallel, with the nitrogen-nitrogen bond being formed from reduced intermediates.  相似文献   

6.
M Tsubaki  T Mogi  H Hori 《FEBS letters》1999,449(2-3):191-195
Azide-binding to the heme-copper binuclear center of bo-type ubiquinol oxidase from Escherichia coli was investigated with Fourier-transform infrared spectroscopy. Deconvolution analyses of infrared spectra of the azide (14N3)-inhibited air-oxidized form showed a major infrared azide antisymmetric stretching band at 2041 cm(-1). An additional band developed at 2062.5 cm(-1) during a longer incubation. Isotope substitutions with terminally 15N-labelled azides did not show a splitting of the major band, indicating that the geometry of the bound azide is mainly in a bridging configuration between high-spin heme o and CuB. The band at 2062.5 cm(-1) showed clear splittings upon substitution with the terminally 15N-labelled azides, indicating the Cu(2+)B-N=N=N structure. Partial reduction of the oxidase with beta-NADH in the presence of azide caused an appearance of new infrared bands at 2038.5 (major) and 2009 (minor) cm(-1). The former band also showed clear splittings in the presence of the terminally 15N-labelled azides, indicating that reduction of low-spin heme b alters the structure of the binuclear center leading to the Fe(3+)o-N=N=N configuration.  相似文献   

7.
The near ubiquitous distribution of nifH genes in sediments sampled from 14 high-temperature (48.0-89.0°C) and acidic (pH 1.90-5.02) geothermal springs in Yellowstone National Park suggested a role for the biological reduction of dinitrogen (N(2)) to ammonia (NH(3)) (e.g. nitrogen fixation or diazotrophy) in these environments. nifH genes from these environments formed three unique phylotypes that were distantly related to acidiphilic, mesophilic diazotrophs. Acetylene reduction assays and (15) N(2) tracer studies in microcosms containing sediments sampled from acidic and high-temperature environments where nifH genes were detected confirmed the potential for biological N(2) reduction in these environments. Rates of acetylene reduction by sediment-associated populations were positively correlated with the concentration of NH(4)(+), suggesting a potential relationship between NH(4)(+) consumption and N(2) fixation activity. Amendment of microcosms with NH(4)(+) resulted in increased lag times in acetylene reduction assays. Manipulation of incubation temperature and pH in acetylene reduction assays indicated that diazotrophic populations are specifically adapted to local conditions. Incubation of sediments in the presence of a N(2) headspace yielded a highly enriched culture containing a single nifH phylotype. This phylotype was detected in all 14 geothermal spring sediments examined and its abundance ranged from ≈ 780 to ≈ 6800 copies (g dry weight sediment)(-1), suggesting that this organism may contribute N to the ecosystems. Collectively, these results for the first time demonstrate thermoacidiphilic N(2) fixation in the natural environment and extend the upper temperature for biological N(2) fixation in terrestrial systems.  相似文献   

8.
NH4(+)-transport in Anabaena 7120 was studied using the NH4+ analogue, 14CH3NH3+. At pH 7, two energy-dependent NH4(+)-transport systems were detected in both N2- and NO3(-)-grown cells, but none in NH4(+)-grown cells. Both transport systems showed a low and a high affinity mode of operation depending on the substrate concentration. One of the transport systems showed Km values of 8 microM (Vmax = 1 nmole min-1mg-1protein) and 80 microM (Vmax = 7 nmole min-1mg-1protein), and was insensitive to L-methionine-DL-sulphoximine, a glutamate analogue and irreversible inhibitor of glutamine synthetase. The other transport system showed Km values of 2.5 microM (Vmax = 0.1 nmole min-1mg-1protein) and 70 microM (Vmax = 0.7 nmole min-1mg-1protein), and was sensitive to L-methionine-DL-sulphoximine. Intracellular accumulation of free 14CH3NH3+ showed a biphasic pattern in response to variation in external 14CH3NH3+ concentrations. A maximum intracellular concentration of 2.5 mM and 7.5 mM was reached in the external 14CH3NH3+ concentration range of 1-50 microM and 1-500 microM, respectively. At pH 9, an energy-independent diffusion of 14CH3NH2 leading to a higher intracellular accumulation and assimilation rate, than that at pH 7, was observed.  相似文献   

9.
Fisher K  Dilworth MJ  Newton WE 《Biochemistry》2000,39(50):15570-15577
In contrast to the wild-type MoFe protein, neither the alpha-195(Asn) nor the alpha-191(Lys) MoFe protein catalyzed N(2) reduction to NH(3), when complemented with wild-type Fe protein. However, N(2) was bound by the alpha-195(Asn) MoFe protein and inhibited the reduction of both protons and C(2)H(2). The alpha-191(Lys) MoFe protein did not interact with N(2). With the alpha-195(Asn) MoFe protein, the N(2)-induced inhibition of substrate reduction was reversed by removing the N(2). Surprisingly, even though added H(2) also relieved N(2) inhibition of substrate reduction, the alpha-195(Asn) MoFe protein did not catalyze HD formation under a N(2)/D(2) atmosphere. This observation is the first indication that these two reactions have different chemical origins, prompting a revision of the current hypothesis that these two reactions are consequences of the same nitrogenase chemistry. A rationale that accounts for the dichotomy of the two reactions is presented. The two altered MoFe proteins also responded quite differently to azide. It was a poor substrate for both but, in addition, azide was an electron-flux inhibitor with the 195(Asn) MoFe protein. The observed reactivity changes are correlated with likely structural changes caused by the amino acid substitutions and provide important details about the interaction(s) of N(2,) H(2), D(2), and azide with Mo-nitrogenase.  相似文献   

10.
The NH(4)(+) and NO(3)(-) uptake kinetics by Typha latifolia L. were studied after prolonged hydroponics growth at constant pH 3.5, 5.0, 6.5 or 7.0 and with NH(4)(+) or NO(3)(-) as the sole N-source. In addition, the effects of pH and N source on H(+) extrusion and adenine nucleotide content were examined. Typha latifolia was able to grow with both N sources at near neutral pH levels, but the plants had higher relative growth rates, higher tissue concentrations of the major nutrients, higher contents of adenine nucleotides, and higher affinity for uptake of inorganic nitrogen when grown on NH(4)(+). Growth almost completely stopped at pH 3.5, irrespective of N source, probably as a consequence of pH effects on plasma membrane integrity and H(+) influx into the root cells. Tissue concentrations of the major nutrients and adenine nucleotides were severely reduced at low pH, and the uptake capacity for inorganic nitrogen was low, and more so for NO(3)(-)-fed than for NH(4)(+)-fed plants. The maximum uptake rate, V(max), was highest for NH(4)(+) at pH 6.5 (30.9 micro mol h(-1) g(-1) root dry weight) and for NO(3)(-) at pH 5.0 (31.7 micro mol h(-1) g(-1) root dry weight), and less than 10% of these values at pH 3.5. The affinity for uptake as estimated by the half saturation constant, K((1/2)), was lowest at low pH for NH(4)(+) and at high pH for NO(3)(-). The changes in V(max) and K((1/2)) were thus consistent with the theory of increasing competition between cations and H(+) at low pH and between anions and OH(-) at high pH. C(min) was independent of pH, but slightly higher for NO(3)(-) than for NH(4)(+) (C(min)(NH(4)(+)) approximately 0.8 mmol m(-3); C(min)(NO(3)(-)) approximately 2.8 mmol m(-3)). The growth inhibition at low pH was probably due to a reduced nutrient uptake and a consequential limitation of growth by nutrient stress. Typha latifolia seems to be well adapted to growth in wetland soils where NH(4)(+) is the prevailing nitrogen compound, but very low pH levels around the roots are very stressful for the plant. The common occurrence of T. latifolia in very acidic areas is probably only possible because of the plant's ability to modify pH-conditions in the rhizosphere.  相似文献   

11.
Proton NMR studies are reported on the complementary d(C-A-T-G-G-G-T-A-C).d(G-T-A-C-epsilon A-C-A-T-G) nonanucleotide duplex (designated epsilon dA.dG 9-mer duplex), which contains exocyclic adduct 1,N6-ethenodeoxyadenosine positioned opposite deoxyguanosine in the center of the helix. The present study focuses on the alignment of dG5 and epsilon dA14 at the lesion site in the epsilon dA.dG 9-mer duplex at neutral pH. This alignment has been characterized by monitoring the NOEs originating from the NH1 proton of dG5 and the H2, H5, and H7/H8 protons of epsilon dA14 in the central d(G4-G5-G6).d(C13-epsilon A14-C15) trinucleotide segment of the epsilon dA.dG 9-mer duplex. These NOE patterns establish that epsilon dA14 adopts a syn glycosidic torsion angle that positions the exocyclic ring toward the major groove edge while all the other bases including dG5 adopt anti glycosidic torsion angles. We detect a set of intra- and interstrand NOEs between protons (exchangeable and nonexchangeable) on adjacent residues in the d(G4-G5-G6).d(C13-epsilon A14-C15) trinucleotide segment which establish formation of right-handed helical conformations on both strands and stacking of the dG5(anti).epsilon dA14(syn) pair between stable dG4.dC15 and dG6.dC13 pairs. The energy-minimized conformation of the central d(G4-G5-G6).d(C13-epsilon A14-C15) segment establishes that the dG5(anti).epsilon dA14(syn) alignment is stabilized by two hydrogen bonds from the NH1 and NH2-2 of dG5(anti) to N9 and N1 of epsilon dA14(syn), respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Nitrogenase catalyzes the sequential addition of six electrons and six protons to a N2 that is bound to the active site metal cluster FeMo-cofactor, yielding two ammonia molecules. The nature of the intermediates bound to FeMo-cofactor along this reduction pathway remains unknown, although it has been suggested that there are intermediates at the level of reduction of diazene (HN=NH, also called diimide) and hydrazine (H2N-NH2). Through in situ generation of diazene during nitrogenase turnover, we show that diazene is a substrate for the wild-type nitrogenase and is reduced to NH3. Diazene reduction, like N2 reduction, is inhibited by H2. This contrasts with the absence of H2 inhibition when nitrogenase reduces hydrazine. These results support the existence of an intermediate early in the N2 reduction pathway at the level of reduction of diazene. Freeze-quenching a MoFe protein variant with alpha-195His substituted by Gln and alpha-70Val substituted by Ala during steady-state turnover with diazene resulted in conversion of the S = 3/2 resting state FeMo-cofactor to a novel S = 1/2 state with g1 = 2.09, g2 = 2.01, and g3 approximately 1.98. 15N- and 1H-ENDOR establish that this state consists of a diazene-derived [-NHx] moiety bound to FeMo-cofactor. This moiety is indistinguishable from the hydrazine-derived [-NHx] moiety bound to FeMo-cofactor when the same MoFe protein is trapped during turnover with hydrazine. These observations suggest that diazene joins the normal N2-reduction pathway, and that the diazene- and hydrazine-trapped turnover states represent the same intermediate in the normal reduction of N2 by nitrogenase. Implications of these findings for the mechanism of N2 reduction by nitrogenase are discussed.  相似文献   

13.
The substitution reactions of tetrakis-(4-N-methylpyridyl)porphinecobalt (III) (CoIIITMpyP) with azide and with 1,1,3,3-tetramethyl-2-thiourea (TMTU) have been studied as a function of pH at 25 degrees and an ionic strength of 0.5 M. The mechanistic pathway proposed for thiocyanate [1] and pyridine [2] is applicable to these ligands as well once allowance is made for two attacking forms of azide, N3- and HN3. A TMTU axial substituent has about the same influence on the rate of further ligand substitution as does SCN- and a much larger influence than does azide. Similar behavior between bound SCN- and bound TMTU is also shown in electron-transfer reactions with Ru(NH3)62+. Whereas both sulfur-containing ligands enhance the rate relative to the diaquo complex, the azide complex undergoes reduction an order of magnitude more slowly than does the diaquo complex.  相似文献   

14.
The distribution of plasmalemmal V-type H+-pumps (V-ATPase) among mammalian macrophages (mvarphi) is uncertain and, hence, the functional significance of mvarphi plasmalemmal V-ATPase is unclear. This study investigated the role of V-ATPase in the regulation of intracellular pH (pH(i)) by resident alveolar mvarphi from sheep, pigs, dogs and rabbits. The fluorescent probe 2',7'-biscarboxyethyl-5,6-carboxyfluorescein was used to monitor baseline pH(i) and the rate of pH(i) recovery (dpH(i)/dt) from intracellular acid-loads (NH(4)Cl prepulse). Baseline pH(i) was 7.1-7.2. In sheep, pig and dog studies, 10 microM bafilomycin A(1) (a selective V-ATPase inhibitor) caused a rapid fall in baseline pH(i) (0.15-0.20 units); baseline values were unaffected by 0.1 mM amiloride (a Na+ transport inhibitor). V-ATPase activity (bafilomycin-sensitive component of dpH(i)/dt) was solely responsible for pH(i) recovery from intracellular acid-loads at acid-loaded pH(i) values >6.8-6.9. Na+/H+ exchange (amiloride-sensitive component of dpH(i)/dt) was detected only at acid-loaded pH(i) values <6.8. The activity of both H+ extruders increased at lower pH(i) values, albeit the Na+/H+ exchanger was more pH-sensitive than was V-ATPase. In rabbit studies, 10 microM bafilomycin A(1) and 1 mM N-ethylmaleimide (a non-specific H+-pump inhibitor) produced similar falls in baseline mvarphi pH(i), but had significantly larger effects than did the selective V-ATPase inhibitor concanamycin A (相似文献   

15.
Hydrazinonicotinamide (HYNIC) forms stable coordination complexes with Tc-99m when reacted with Tc(V)oxo species such as Tc-mannitol or other Tc-polyhydric complexes. However, radio-HPLC of [Tc-For-MLFK-HYNIC] labeled via Tc-polyhydric ligands demonstrated multiple radiochemical species each with unique biodistribution patterns. This is likely due to the fact that Tc can bind to the hydrazino moiety, as well as polyhydric ligands, in a variety of coordination geometries. Tridentate ligands, such as bis(mercaptoethyl)methylamine (NS2), may constrain the possible coordination geometries and improve overall stability. To investigate this, we synthesized NS2, converted the [Tc-mannitol-For-MLFK-HYNIC] to the corresponding NS2-containing complex [Tc-NS2-For-MLFK-HYNIC], and compared its infection imaging and biodistribution properties with [Tc-mannitol-For-MLFK-HYNIC]. Conversion to the NS2 complex was confirmed by HPLC which showed a single unique hydrophobic species with retention time greater than the [Tc-mannitol-For-MLFK-HYNIC] complex. Imaging experiments with both preparations were performed in rabbits with E. coli infections in the left thigh. Tissue radioactivity measurements demonstrated that compared to Tc-mannitol-peptide, accumulation of Tc-NS2-peptide was lower in blood, heart, and normal muscle and higher in spleen, infected muscle, and pus (p < 0.01). These results indicate that the Tc-NS2-peptide complex is chemically more homogeneous and exhibits improved infection localization and biodistribution properties. In an effort to model the interactions of the metal-HYNIC core with NS2 and related ligand types, the reactions of [ReCl3(NNC5H4NH)(NHNC5H4N)] and [99TcCl3(NNC5H4NH)(NHNC5H4N)], effective structural analogues for the [M(NNC5H4NH(x))2] core, with NS2, C5H3N-2,6-(CH2SH)2, O(CH2CH2SH)2, and S(CH2CH2SH)2 were investigated and the compounds [M[CH3N(CH2CH2S)2](NNC5H4N)(NHNC5H4N] (M = 99Tc (5a), Re (5b)), [Re[C5H3N-2,6-(CH2S)2](NNC5H4N)(NHNC5H4N)].CH2Cl2.0.5MeOH (7), [Re[SCH2CH2)2O] (NNC5H4N)(NHNC5H4N)] (8), and [Re[(SCH2CH2)2S](NNC5H4NH)(NHNC5H4N)]Cl (9) were isolated. Similarly, the reaction of [ReCl3(NNC5H4NH)(NHNC5H4N)] with the bidentate ligands pyridine-2-methanethiol and 3-(trimethlysilyl)pyridine-2-thiol led to the isolation of [ReCl(C5H4N-2-CH2S) (NNC5H4N)(NHNC5H4N)] (10) and [Re(2-SC5H3N-3-SiMe3)2 (NNC5H4N)(NHNC5H4N)] (11), respectively, while reaction with N-methylimidazole-2-thiol yielded the binuclear complex [Re(OH)Cl(SC3H2N2CH3)2(NNC5H4N)2 (NHNC5H4N)2] (12). The analogous metal-(HYNIC-OH) precursor, [ReCl3[NNC5H3NH(CO2R)] [NHNC5H3N(CO2R)]] (R = H, 13a; R = CH3, 13b) has been prepared and coupled to lysine to provide [RCl3[NNC5H3NH(CONHCH2CH2CH2CH2CH(NH2)CO2H)] [NHNC5H3NH(CONHCH2CH2CH2CH2CH(NH2)CO2H)]].2HCl (14.2HCl), while the reaction of the methyl ester 13b with 2-mercaptopyridine yields [Re(2-SC5H4N)2[NNC5H3N(CO2Me)][NHNC5H3N(CO2Me)]] (15). While the chemical studies confirm the robustness of the M-HYNIC core (M = Tc, Re) and its persistence in ligand substitution reactions at adjacent coordination sites of the metal, the isolation of oligomeric structures and the insolubility of the peptide conjugates of 13, 14, and 15 underscore the difficulty of characterizing these materials on the macroscopic scale, an observation relevant to the persistent concerns with reagent purity and identity on the tracer level.  相似文献   

16.
The kinetics of the reaction of Golfingia gouldii hemerythrin with O2 have been studied by stopped flow spectrophotometry. For the second order oxygenation process, k1 = 7.4 X 10(6) M-1 s-1, deltaH1++ = 8.2 kcal-mol-1 and deltaS1++ = +1 e.u. at 25 degrees, pH 8.2, and I = 0.015 M. The rate constant is unchanged when protein concentration is varied from 3 to 25 muM, the ionic strength is increased to 0.07 M, and the pH moved to 6.8. The deoxygenation of oxyhemerythrin is studied with stopped flow by scavenging liberated O2 with S2O4(2-). For the first order dissociation, k-1 = 51 s-1, deltaH-1++ = 20.6 kcal-mol-1 and deltaS-1++ = +19 e.u. at 25 degrees, pH 8.2, and I = 0.015 M. The value of k-1 is independent of [protein] = 50 to 200 muM, [S2O4(2-)] = 5 to 100 mM I = 0.015 to 0.30 M and pH 6.8 to 9.0. Using myoglobin instead of S2O4(2-) as scavenger gives similar results. Combination of activation parameters for the oxygenation and deoxygenation processes gives K1 = 1.5 X 10(5) M-1, deltaH = -12.4 kcal-mol-1, and deltaS = -18 e.u., values in good agreement with independent thermodynamic data. Perchlorate ion (0.05 M) enhances k-1 about 3-fold and hardly effects k1. There is no sign of other than a single reaction in either direction, and octameric hemerythrin apparently behaves kinetically as eight single units.  相似文献   

17.
Exchange-inert beta, gamma-bidentate Cr(H2O)x(NH3)y ATP complexes inactivate yeast phosphoglycerate kinase (PGK) by forming a coordination complex at the enzyme active site. The observed inactivation rates ranged from 0.019 min-1 to 0.118 min-1 for Cr(NH3)4ATP and Cr(H2O)4ATP, respectively. Incorporation of one mol of Cr-ATP to the enzyme was sufficient for complete inactivation of the enzyme. The presence of Mg-ATP protected the enzyme against inactivation by Cr-ATP. The other substrate 3-phosphoglycerate (3-PGA), when present, reduced the observed inactivation rates. The reduction of the k(obs) by 3-PGA was proportional to the number of NH3 ligands present in the coordination sphere of Cr3+ in the Cr-ATP complex, suggesting that in the ternary enzyme-Cr-ATP-3-PGA complex 3-PGA may be coordinated to the metal ion. When the effector sulfate ion was present, the presence of 3-PGA did not cause any further effects on the observed inactivation rates. This suggests that bound substrates are in a different arrangement at the active site when sulfate is present and therefore 3-PGA may not need to displace a ligand from Cr3+. Additionally, PGK exhibited a stereoselectivity for the binding of Cr(H2O)4ATP. delta diastereomer of Cr(H2O)4ATP yielded an order of magnitude smaller Ki value compared to the value observed with the lambda isomer. The recovery of enzyme activity was observed over a period of a few hours upon removal of excess Cr-ATP. The presence of substrates and/or effector ion sulfate did not alter the observed reactivation rate. There was no difference in the reactivation rates of the enzyme which was inactivated with Cr(H2O)4ATP or Cr(NH3)4ATP with and without 3-PGA. Increasing the ligand exchange rates of Cr3+ of Cr-ATP by increasing the pH value of the recovery medium from 5.9 to 6.8 increased the rate of recovery by a factor of 8. The pH dependence of the reactivation indicated that one hydroxyl group is involved in the recovery of the enzyme activity in enzyme CrATP and enzyme.CrATP.3-PGA complexes.  相似文献   

18.
Ammonium isosaccharate, C6H15NO6.H2O (NH4-ISA), has been synthesized and its crystal structure solved by single-crystal X-ray diffraction methods. NH4-ISA crystallizes in the monoclinic space group P2(1) (#4) with cell parameters a=8.6470(12)A, b=5.0207(7)A, c=9.8193(14)A, beta=91.643(3) degrees , V=426.12(10)A3, Z=2. The structure was refined by full-matrix least-squares on F2 yielding final R-values (all data) R1=0.0485 and Rw2=0.1104. The structure consists of alternating (NH4)+ and (C6H11O6)- layers parallel to the ab plane. An extended network of O-H...O intermolecular (ISA)...(ISA) hydrogen bonds links the (ISA)- anions within the ab plane, while the 3-D connectivity along the c-axis is provided only by (ISA-)...(NH4+)...(ISA-) hydrogen bonds. The aqueous solubility (Si, [ML(-1)]) of NH4- and Na-ISA has been shown to be pH independent at ambient conditions within the range 4.5相似文献   

19.
The oxidation of cytochromes during the reduction of N2O to N2 by a denitrifying bacterium was studied spectrophotometrically. The reduced b- and c-type cytochromes are partially oxidized when N2O is added to intact cells reduced with lactate under anaerobic conditions. The oxidation of cytochromes is inhibited non-competitively by azide, cyanide, 2,4-dinitrophenol and CuSO4, which inhibit the reduction of N2O to N2. In the presence of each inhibitor at a high concentration, at which the reduction of N2O to N2 is perfectly inhibited, cytochromes are not oxidized by N2O, while when an adequate, low concentration of inhibitor is added, b-type cytochrome is partially oxidized but c-type cytochrome is apparently not oxidized. In cell-free extracts, prepared by the sonic disruption of cells, that have entirely lost their activity in the reduction of N2O to N2, cytochromes are not oxidized by N2O. From the above results, it was concluded that b-type and c-type cytochromes should participate in the electron transport mechanism of the reduction of N2O to N2.  相似文献   

20.
Seven new antagonists of bombesin (Bn)/gastrin-releasing peptide (GRP) containing C-terminal Trp or Tpi (2,3,4,9-tetrahydro-1H-pyrido[3,4-b]indol-3-carboxylic acid) in a reduced peptide bond were synthesized by solid phase methods and evaluated biologically. The reduced bond in four [Leu13 psi(CH2NH)Trp14]Bn(6-14) analogs was formed by reductive alkylation at the dipeptide stage. In the case of three [Leu13 psi(CH2N)Tpi14]Bn(6-14) analogs, the Trp dipeptide with reduced bond was reacted with formaldehyde to form the corresponding Tpi derivative. These Tpi-containing analogs have a new reduced bond which is structurally more constrained. Leu13 psi(CH2N)Tpi14 analogs inhibit [125I][Tyr4]bombesin binding to Swiss 3T3 cells with IC50 values of 2-4 nM, compared to 5-10 nM for Leu13 psi(CH2NH)Trp14 analogs. Leu13 psi(CH2N)Tpi14 analogs are also more potent than Leu13 psi(CH2NH)Trp14 analogs in growth inhibition studies using Swiss 3T3 cells. The two best bombesin antagonists of this series, [D-Trp6,Leu13 psi(CH2N)Tpi14]Bn(6-14) (RC-3415) and [Tpi6,Leu13 psi(CH2N)Tpi14]Bn(6-14) (RC-3440), inhibited GRP-stimulated growth of Swiss 3T3 cells with IC50 values less than 1 nM. RC-3440 was also active in vivo, suppressing GRP(14-27)-stimulated serum gastrin secretion in rats. Bombesin/GRP antagonists, such as RC-3440, containing the new reduced bond (CH2N) reported herein are very potent.  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