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.
Azotobacter vinelandii DJ71, which contains a mutation in the nifV gene, was derepressed for nitrogenase in the presence of homocitrate. When dinitrogenase was isolated from this culture, it was found to be identical to the wild-type dinitrogenase. However, when the same NifV- strain was derepressed in the presence of erythrofluorohomocitrate, a homocitrate analog which produces a nitrogenase with wild-type properties in vitro, the isolated dinitrogenase was characteristic of the NifV- enzyme. These data show that homocitrate, but not fluorohomocitrate, is utilized by NifV- mutant cells. Fluorohomocitrate does not inhibit the uptake of homocitrate because the wild-type phenotype resulted when both compounds were added to the medium during nitrogenase derepression. Homocitrate lactone failed to cure the NifV- phenotype.  相似文献   

2.
The in vitro synthesis of the iron-molybdenum cofactor (FeMo-co) of nitrogenase requires homocitrate (2-hydroxy-1,2,4-butanetricarboxylic acid). Homocitrate is apparently synthesized by the nifV gene product. In the absence of homocitrate, no FeMo-co is formed in vitro, as determined from coupled C2H2 reduction assays and the lack of 99Mo label incorporation into apodinitrogenase. Several organic acids were tested for their ability to replace homocitrate in the FeMo-co synthesis system. With appropriate homocitrate analogues, aberrant forms of FeMo-co are synthesized that exhibit altered substrate specificity and inhibitor susceptibility. Homoisocitrate (1-hydroxy-1,2,4-butanetricarboxylic acid) and 2-oxoglutarate facilitated the incorporation of 99Mo into apodinitrogenase in the FeMo-co synthesis system, yielding a dinitrogenase that effectively catalyzed the reduction of protons but not C2H2 or N2. Citrate also promoted the incorporation of 99Mo into apodinitrogenase, and the resulting holodinitrogenase reduced protons and C2H2 effectively but not N2. In addition, proton reduction from this enzyme was inhibited by CO. The properties of the homodinitrogenase formed in the presence of citrate were reminiscent of those of the Klebsiella pneumoniae NifV- dinitrogenase. We also observed low rates of HD formation from NifV- dinitrogenase compared to those from the wild-type enzyme. No HD formation was observed with the dinitrogenase activated in vitro in the presence of citrate. We propose that in vivo NifV- mutants utilize citrate for FeMo-co synthesis.  相似文献   

3.
Biosynthesis of the iron-molybdenum cofactor of nitrogenase   总被引:3,自引:0,他引:3  
The iron-molybdenum cofactor (FeMo-co) of nitrogenase is a Mo-Fe-S cluster that has been proposed as the site of substrate reduction for the nitrogenase enzyme complex. Biosynthesis of FeMo-co in Klebsiella pneumoniae requires at least six nif (nitrogen fixation) gene products. One of the nif genes, nifV, apparently encodes a homocitrate synthase. The synthesis and accumulation of homocitrate [(R)-2-hydroxy-1,2,4-butanetricarboxylic acid] in K.pneumoniae is correlated to the presence of a functional nifV gene. K.pneumoniae strains with mutations in nifV synthesize and accumulate an aberrant form of FeMo-co. Nitrogenase from NifV- mutants is capable of reducing some of the substrates of nitrogenase effectively (e.g. acetylene), but reduces N2 poorly. With the aid of an in vitro FeMo-co synthesis system, it recently has been established that homocitrate is an endogenous component of FeMo-co. Substitution of homocitrate with other carboxylic acids results in the formation of aberrant forms of FeMo-co with altered substrate reduction capability.  相似文献   

4.
Mutants of Pichia guilliermondii were isolated that lacked homoaconitate hydratase (lys4), homoisocitrate dehydrogenase (lys10) or α-aminoadipate reductase (lys2) and were able to excrete homocitrate into the culture medium. The effects of incubation time and lysine concentration in the medium on the excretion of homocitrate were examined. In the presence of 600 mg of L-lysine/1 in a minimum salt medium P. guilliermondii G75 (lys2) produced about 280 mg homocitrate/1 during 48 h of growth. A simple procedure to isolate homocitrate from the medium is described.  相似文献   

5.
J Liang  M Madden  V K Shah  R H Burris 《Biochemistry》1990,29(37):8577-8581
An organic acid extracted from purified dinitrogenase isolated from a nifV mutant of Klebsiella pneumoniae has been identified as citric acid. H2 evolution by the citrate-containing dinitrogenase is partially inhibited by CO, and by some substrates for nitrogenase. The response of maximum velocities to changes in pH for both the wild-type and the NifV- dinitrogenase was compared. No substantial differences between the enzymes were observed, but there are minor differences. Both enzymes are stable in the pH range 4.8-10, but the enzyme activities dropped dramatically below pH 6.2.  相似文献   

6.
7.
钼铁蛋白铁钼辅因子的有机组分对其功能的影响   总被引:3,自引:0,他引:3  
棕色固氮菌(Azotobacter vinelandii)固氮酶的钼铁蛋白经邻菲啰啉在厌氧或有氧环境中处理后,变为 P-cluster 单一缺失或 P-cluster 和 FeMoco 同时缺失的失活钼铁蛋白。含柠檬酸盐或高柠檬酸盐的重组液都使这两种失活蛋白能恢复固氮酶重组的 H~ 和 C_2H_2还原活性,活性恢复程度随反映钼铁蛋白中金属原子簇含量变化的圆二色和磁圆二色谱及金属含量的恢复程度的提高而提高,但它们固 N_2能力的恢复程度则不相同:P-cluster 单一缺失的蛋白用两种重组液重组后均可恢复其固 N_2能力,而 P-cluster 和 FeMoco 同时缺失的蛋白,只有用含高柠檬酸盐的重组液重组才恢复其固 N_2能力,表明含不同有机组分的重组液所组装的 P-cluster 均与天然状态相同,只有含高柠檬酸盐的重组液所组装的 FeMoco 才与天然状态相同,从而证明高柠檬酸盐是 FeMoco 的必需的有机组分。  相似文献   

8.
Homocitrate is a component of the iron-molybdenum cofactor of nitrogenase   总被引:18,自引:0,他引:18  
When apodinitrogenase (lacking FeMo-co) was activated with FeMo-co synthesized in vitro in the presence of 3H-labeled homocitrate, label was incorporated into dinitrogenase. The physical association of the label with FeMo-co was demonstrated by reisolation and purification of the cofactor from dinitrogenase. The presence of homocitrate in FeMo-co was established by NMR analysis of the organic acid extracted from dinitrogenase. Quantitation of homocitrate in dinitrogenase showed it to be present at a 1:1 ratio with molybdenum.  相似文献   

9.
Summary Eight strains devoid of homocitrate synthase activity were found among lysine requiring mutants of the yeast Saccharomycopsis lipolytica. Genetic analysis of these strains showed that they were all affected at the same locus LYS 1. Three lines of evidence suggest that this locus defines a structural gene for homocitrate synthase. First, the mutations show various degrees of intragenic complementation; it could be shown in some cases that the hybrid enzyme formed in vivo displayed modified properties in vitro. Second, reversion of some of these mutations can result in a modified enzyme (desensitized). Third, a feedback mutant of homocitrate synthase was directly isolated from the wild type strain, and shown to carry a single mutation at or near LYS 1.We also present here the first attempts at genetic fine mapping in Saccharomycopsis lipolytica.Abbreviations used lys lysine - arg arginine - ade adenine - ura uracile - TDL 4,5-transdehydrolysine - Sm Saccharomycopsis - KR kilorads Part of a thesis submitted by C.G. to the Université de Paris VI, Paris, France  相似文献   

10.
Homocitrate synthase in the first enzyme of the lysine biosynthetic pathway. It is feedback regulated by L-lysine. Lysine decreases the biosynthesis of penicillin (determined by the incorporation of [14C]valine into penicillin) by inhibiting and repressing homocitrate synthase, thereby depriving the cell of alpha-aminoadipic acid, a precursor of penicillin. Lysine feedback inhibited in vivo the biosynthesis and excretion of homocitrate by a lysine auxotroph, L2, blocked in the lysine pathway after homocitrate. Neither penicillin nor 6-aminopenicillanic acid exerted any effect at the homocitrate synthase level. The molecular mechanism of lysine feedback regulation in Penicillium chrysogenum involved both inhibition of homocitrate synthase activity and repression of its synthesis. In vitro studies indicated that L-lysine feedback inhibits and represses homocitrate synthase both in low- and high-penicillin-producing strains. Inhibition of homocitrate synthase activity by lysine was observed in cells in which protein synthesis was arrested with cycloheximide. Maximum homocitrate synthase activity in cultures of P. chrysogenum AS-P-78 was found at 48 h, coinciding with the phase of high rate of penicillin biosynthesis.  相似文献   

11.
There are conflicting reports regarding the cellular localization in Saccharomyces cerevisiae and filamentous fungi of homocitrate synthase, the first enzyme in the lysine biosynthetic pathway. The homocitrate synthase (HS) gene (lys1) of Penicillium chrysogenum was disrupted in three transformants (HS(-)) of the Wis 54-1255 pyrG strain. The three mutants named HS1(-), HS2(-) and HS3(-) all lacked homocitrate synthase activity and showed lysine auxotrophy, indicating that there is a single gene for homocitrate synthase in P. chrysogenum. The lys1 ORF was fused in frame to the gene for the green fluorescent protein (GFP) gene of the jellyfish Aequorea victoria. Homocitrate synthase-deficient mutants transformed with a plasmid containing the lys1-GFP fusion recovered prototrophy and showed similar levels of homocitrate synthase activity to the parental strain Wis 54-1255, indicating that the hybrid protein retains the biological function of wild-type homocitrate synthase. Immunoblotting analysis revealed that the HS-GFP fusion protein is maintained intact and does not release the GFP moiety. Fluorescence microscopy analysis of the transformants showed that homocitrate synthase was mainly located in the cytoplasm in P. chrysogenum; in S. cerevisiae the enzyme is targeted to the nucleus. The control nuclear protein StuA was properly targeted to the nucleus when the StuA (targeting domain)-GFP hybrid protein was expressed in P. chrysogenum. The difference in localization of homocitrate synthase between P. chrysogenum and S. cerevisiae suggests that this protein may play a regulatory function, in addition to its catalytic function, in S. cerevisiae but not in P. chrysogenum.  相似文献   

12.
Nitrogenase of Klebsiella pneumoniae nifV mutants.   总被引:7,自引:2,他引:5       下载免费PDF全文
The MoFe protein of nitrogenase from Klebsiella pneumoniae nifV mutants, NifV- Kp1 protein, in combination with the Fe protein from wild-type cells, catalysed CO-sensitive H2 evolution, in contrast with the CO-insensitive reaction catalysed by the wild-type enzyme. The decrease in H2 production was accompanied by a stoicheiometric decrease in dithionite (reductant) utilization, implying that CO was not reduced. However, CO did not affect the rate of phosphate release from ATP. Therefore the ATP/2e ratio increased, indicating futile cycling of electrons between the Fe protein and the MoFe protein. The inhibition of H2 evolution by CO was partial; it increased from 40% at pH6.3 to 82% at pH 8.6. Inhibition at pH7.4 (maximum 73%) was half-maximal at 3.1 Pa (0.031 matm) CO. The pH optimum of the mutant enzyme was lower in the presence of CO. Steady-state kinetic analysis of acetylene reduction indicated that CO was a linear, intersecting, non-competitive inhibitor of acetylene reduction with Kii = 2.5 Pa and Kis = 9.5 Pa. This may indicate that a single high-affinity CO-binding site in the NifV- Kp1 protein can cause both partial inhibition of H2 evolution and total elimination of acetylene reduction. Various models to explain the data are discussed.  相似文献   

13.
Abstract We isolated from Candida maltosa mutants lacking saccharopine reductase ( lys9 ) and saccharopine dehydrogenase ( lys1 ). They accumulated α-aminoadipate-δ-semialdehyde (AASA) in the cell and excreted it into the culture medium. In the presence of 15 g glucose/l, 1.25 g NH4H2PO4/l and 50 mg l -lysine/l in a minimal salt medium C. maltosa G285 ( lys1 ) produced about 80–90 mg AASA/l within 48 h. It is the first report of lysine-requiring yeast mutants that accumulate and excrete AASA. In contrast, Pichia guilliermondii lys9 mutants lacked this AASA overproduction. The AASA accumulation by C. maltosa mutants may be explained by the low feedback regulation of their homocitrate synthase and the equilibrium of the enzyme reactions involved in the lysine biosynthesis.  相似文献   

14.
We previously reported that lysine inhibits in vivo homocitrate synthesis in the lysine bradytroph, Penicillium chrysogenum L(1), and that such feedback inhibition could explain the known lysine inhibition of penicillin formation. In the present study, it was found that dialyzed cell-free extracts of mutant L(1) converted [1-(14)C]acetate to homocitrate. This homocitrate synthase activity was extremely labile but could be stabilized by high salt concentrations. The pH optimum of the reaction was 6.9, and the K(m) was 5.5 mM with respect to alpha-ketoglutarate. The reaction was also dependent upon the presence of Mg(2+), adenosine 5'-triphosphate, and coenzyme A. Surprisingly, the activity in these crude extracts was not inhibited by lysine. Benzylpenicillin at a high concentration (20 mM) partially inhibited the enzyme, an effect that was enhanced by lysine. Casein hydrolysate also partially inhibited the enzyme.  相似文献   

15.
Andi B  West AH  Cook PF 《Biochemistry》2004,43(37):11790-11795
Kinetic data have been collected suggesting a preferred sequential ordered kinetic mechanism for the histidine-tagged homocitrate synthase (HCS) from Saccharomyces cerevisiae with alpha-ketoglutarate binding before AcCoA and CoA released before homocitrate. Oxaloacetate is also a substrate for HCS, but with lower affinity than alpha-ketoglutarate. In agreement with the ordered kinetic mechanism desulfo-CoA is uncompetitive and citrate is competitive vs alpha-ketoglutarate. Varying AcCoA, citrate is a noncompetitive inhibitor as predicted, but CoA is noncompetitive vs AcCoA suggesting binding of CoA to E:homocitrate and E:alpha-ketoglutarate. The product CoA behaves in a manner identical to the dead-end analogue desulfo-CoA, suggesting an E:alpha-ketoglutarate:CoA dead-end complex. Data further suggest an irreversible reaction overall, in agreement with the downhill nature of the reaction as a result of homocitryl-CoA hydrolysis. Fluorescence titration data generally agree with the steady state data, but show finite binding of CoA and AcCoA to free enzyme, suggesting that the mechanism may be random with a high degree of synergism of binding between the reactants.  相似文献   

16.
(R)-2-Hydroxy-1,2,4-butanetricarboxylic acid [(R)-homocitrate] has been has been recently reported to be an integral constituent of the otherwise thought to be inorganic iron-molybdenum cofactor of dinitrogenase [Hoover, T.R., Imperial, J., Ludden, P.W., & Shah, V.K. (1989) Biochemistry 28,2768-2771]. Different organic acids can substitute for homocitrate in an in vitro system for iron-molybdenum cofactor synthesis and incorporation into dinitrogenase [Hoover, T.R., Imperial, J., Ludden, P.W., & Shah, V. K. (1988) Biochemistry 27, 3647-3652]. Dinitrogenase activated with homocitrate-FeMo-co was able to reduce dinitrogen, acetylene, and protons efficiently. Homoisocitrate and isocitrate dinitrogenases did not reduce dinitrogen or acetylene, but showed very high proton reduction activities. Citrate and citramalate dinitrogenases had very low dinitrogen reduction activities and intermediate acetylene and proton reduction activities. CO inhibited proton reduction in both these cases but not in the case of dinitrogenases activated with other homocitrate analogues. By use of these and other commercially available homocitrate analogues in the in vitro system, the structural features of the homocitrate molecule absolutely required for the synthesis of a catalytically competent iron-molybdenum cofactor were determined to be the hydroxyl group, the 1- and 2-carboxyl groups, and the R configuration of the chiral center. The stringency of the structural requirements was dependent on the nitrogenase substrate used for the assay, with dinitrogen having the most stringent requirements followed by acetylene and protons.  相似文献   

17.
Qian J  West AH  Cook PF 《Biochemistry》2006,45(39):12136-12143
Homocitrate synthase (acetyl-coenzyme A:2-ketoglutarate C-transferase; E.C. 2.3.3.14) catalyzes the condensation of AcCoA and alpha-ketoglutarate to give homocitrate and CoA. The enzyme was found to be a Zn-containing metalloenzyme using inductively coupled plasma mass spectrometry. Dead-end analogues of alpha-ketoglutarate were used to obtain information on the topography of the alpha-ketoglutarate binding site. The alpha-carboxylate and alpha-oxo groups of alpha-ketoglutarate are required for optimum binding to coordinate to the active site Zn. Optimum positioning of the alpha-carboxylate, alpha-oxo, and gamma-carboxylate of alpha-ketoglutarate is likely mimicked by the location in space of the 2-carboxylate, pyridine nitrogen, and 4 carboxylate of pyridine 2,4-dicarboxylate. The pH dependence of the kinetic parameters was determined to obtain information on the chemical mechanism of homocitrate synthase. The V profile is bell shaped with slopes of 1 and -1, giving pKa values of 6.7 and 8.0, while V/K(AcCoA) exhibits a slope of 2 on the acidic side with an average pKa value of 6.6 and a slope of -2 on basic side of the profile with an average pKa value of 8.2. The V/K(alpha-Kg) pH-rate profile exhibits a single pKa of 6.9 on the acidic side and two on the basic side with an average value of 7.8. The pH dependence of the Ki for glyoxylate, a competitive inhibitor vs alpha-ketoglutarate, gives a pKa of 7.1 for a group, required to be protonated for optimum binding. Data suggest a chemical mechanism for the enzyme in which alpha-ketoglutarate first binds to the active site Zn via its alpha-carboxylate and alpha-oxo groups, followed by acetyl-CoA. A general base then accepts a proton from the methyl of acetyl-CoA, and a general acid protonates the carbonyl of alpha-ketoglutarate in the formation of homocitryl-CoA. The general acid then acts as a base in deprotonating Zn-OH2 in the hydrolysis of homocitryl-CoA to give homocitrate and CoA. A solvent deuterium kinetic isotope effect of 1 is measured for homocitrate synthase, while a small pH-independent primary kinetic deuterium isotope effect (approximately 1.3) is observed using deuterioacetyl-CoA. Data suggest rate-limiting condensation to form the alkoxide of homocitryl-CoA, followed by hydrolysis to give products.  相似文献   

18.
19.
In the case of nitrogenase-based photobiological hydrogen production systems of cyanobacteria, the inactivation of uptake hydrogenase (Hup) leads to significant increases in hydrogen production activity. However, the high-level-activity stage of the Hup mutants lasts only a few tens of hours under air, a circumstance which seems to be caused by sufficient amounts of combined nitrogen supplied by active nitrogenase. The catalytic FeMo cofactor of nitrogenase binds homocitrate, which is required for efficient nitrogen fixation. It was reported previously that the nitrogenase from the homocitrate synthase gene (nifV) disruption mutant of Klebsiella pneumoniae shows decreased nitrogen fixation activity and increased hydrogen production activity under N2. The cyanobacterium Nostoc sp. strain PCC 7120 has two homocitrate synthase genes, nifV1 and nifV2, and with the ΔhupL variant of Nostoc sp. strain PCC 7120 as the parental strain, we have constructed two single mutants, the ΔhupL ΔnifV1 strain (with the hupL and nifV1 genes disrupted) and the ΔhupL ΔnifV2 strain, and a double mutant, the ΔhupL ΔnifV1 ΔnifV2 strain. Diazotrophic growth rates of the two nifV single mutants and the double mutant were decreased moderately and severely, respectively, compared with the rates of the parent ΔhupL strain. The hydrogen production activity of the ΔhupL ΔnifV1 mutant was sustained at higher levels than the activity of the parent ΔhupL strain after about 2 days of combined-nitrogen step down, and the activity in the culture of the former became higher than that in the culture of the latter. The presence of N2 gas inhibited hydrogen production in the ΔhupL ΔnifV1 ΔnifV2 mutant less strongly than in the parent ΔhupL strain and the ΔhupL ΔnifV1 and ΔhupL ΔnifV2 mutants. The alteration of homocitrate synthase activity can be a useful strategy for improving sustained photobiological hydrogen production in cyanobacteria.  相似文献   

20.
A vanadium- and iron-containing cluster has been shown previously to accumulate on VnfX in the Azotobacter vinelandii mutant strain CA11.1 (DeltanifHDKvnfDGK::spc). In the present study, we show the homocitrate-dependent transfer of (49)V label from VnfX to nif-apodinitrogenase in vitro. This transfer of radiolabel correlates with acquisition of acetylene reduction activity. Acetylene is reduced both to ethylene and ethane by the hybrid holodinitrogenase so formed, a feature characteristic of alternative nitrogenases. Structural analogues of homocitrate prevent the acetylene reduction ability of the resulting dinitrogenase. Addition of NifB cofactor (-co) or a source of vanadium (Na(3)VO(4) or VCl(3)) does not increase nitrogenase activity. Our results suggest that there is in vitro incorporation of homocitrate into the V-Fe-S cluster associated with VnfX and that the completed cluster can be inserted into nif-apodinitrogenase. The homocitrate incorporation reaction and the insertion of the cluster into nif-apodinitrogenase (alpha(2)beta(2)gamma(2)) do not require MgATP. Attempts to achieve FeV-co synthesis using extracts of other FeV-co-negative mutants were unsuccessful, showing that earlier steps in FeV-co synthesis, such as the steps requiring VnfNE or VnfH, do not occur in vitro.  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