共查询到20条相似文献,搜索用时 15 毫秒
1.
Adenylylsulfate reductase (EC 1.8.99.2) isolated from Desulfovibrio vulgaris Miyazaki catalyzes electron transfer from dihydroflavin coenzyme (FADH2, FMNH2, or dihydroriboflavin) to adenylyl sulfate (APS), and catalyzes flavin-mediated oxidation of ferrocytochrome c3 with APS. The reaction with FAD as an electron mediator was markedly stimulated in the presence of menadione. Km of the enzyme was about 0.015 mM for riboflavin and FAD in the presence of menadione. Free flavin coenzyme was found to be the normal cellular constituent. These observations suggested that free flavin coenzyme may be a physiological electron carrier for APS reductase, and the enzyme may be called AMP, sulfite:flavin oxidoreductase. Km (APS) of this enzyme is lower than 1 microM. The enzyme is not inhibited by ATP and GTP, but was inhibited by AMP and sulfite. Its extremely low Km (APS) enables this enzyme to reduce any traces of cytosolic APS which is present only at micromolar concentration, and inhibition by sulfite makes this organism utilize an energetically favorable electron acceptor, sulfite, preferentially over APS which is produced from sulfate at the cost of ATP. 相似文献
2.
Cloning, sequence and expression of the gene encoding the malolactic enzyme from Lactococcus lactis 总被引:1,自引:0,他引:1
Many lactic acid bacteria can carry out malolactic fermentation. This secondary fermentation is mediated by the NAD- and Mn(2+)-dependent malolactic enzyme, which catalyses the decarboxylation of L-malate to L-lactate. The gene we call mleS, coding for malolactic enzyme, was isolated from Lactococcus lactis. The mleS gene consists of one open reading frame capable of coding for a protein with a calculated molecular mass of 59 kDa. The amino acid sequence of the predicted MleS gene product is homologous to the sequences of different malic enzymes. Bacterial and yeast cells expressing the malolactic gene convert L-malate to L-lactate. 相似文献
3.
A hypothetical model for electron transfer complex between cytochrome c3 and the flavodoxin from the sulfate-reducing bacteria Desulfovibrio vulgaris has been proposed, based on electrostatic potential field calculations and NMR data [Stewart, D. E., LeGall, J. , Moura, I., Moura, J. J. G., Peck, H. D., Jr., Xavier, A. V., Weiner, P. K., & Wampler, J. E. (1988) Biochemistry 27, 2444-2450]. This modeled complex relies primarily on the formation of five ion pairs between lysine residues of the cytochrome and acidic residues surrounding the flavin mononucleotide cofactor of the flavodoxin. In this study, the role of several acidic residues of the flavodoxin in the formation of this complex and in electron transfer between these two proteins was evaluated. A total of 17 flavodoxin mutants were studied in which 10 acidic amino acids--Asp62, Asp63, Glu66, Asp69, Asp70, Asp95, Glu99, Asp106, Asp127, and Asp129--had been permanently neutralized either individually or in various combinations by substitution with their amide amino acid equivalent (i.e., asparate to asparagine, glutamate to glutamine) through site-directed mutagenesis. The kinetic data for the transfer of electrons from reduced cytochrome c3 to the various flavodoxin mutants do not conform well to a simple bimolecular mechanism involving the formation of an intermediate electron transfer complex. Instead, a minimal electron transfer mechanism is proposed in which an initial complex is formed that is stabilized by intermolecular electrostatic interactions but is relatively inefficient in terms of electron transfer. This step is followed by a rate-limiting reorganization of that complex leading to efficient electron transfer. The apparent rate of this reorganization step was enhanced by the disruption of the initial electrostatic interactions through the neutralization of certain acidic amino acid residues leading to faster overall observed electron transfer rates at low ionic strengths. Of the five acidic residues involved in ion pairing in the modeled complex proposed by Stewart et al. (1988), the kinetic data strongly implicate Asp62, Glu66, and Asp95 in the formation of the electrostatic interactions that control electron transfer. Less certainty is provided by this study for the involvement of Asp69 and Asp129, although the data do not exclude their participation. It was not possible to determine whether the modeled complex represents the optimal configuration for electron transfer obtained after the reorganization step or actually represents the initial complex. The data do provide evidence for the importance of electrostatic interactions in electron transfer between these two proteins and for the existence of alternative binding modes involving acidic residues on the surface of the flavodoxin other than those proposed in that model. 相似文献
4.
We cloned a 9.4-kb DNA fragment from Streptomyces scabies ATCC 41973 that allows the nonpathogen Streptomyces lividans 66 TK24 to necrotize and colonize potato tuber slices and produce scab-like symptoms on potato minitubers. Deletion analysis demonstrated that activity was conferred by a 1.6-kb DNA region. Sequence analysis of a 2.4-kb DNA fragment spanning the DNA region necessary for activity revealed three open reading frames (ORFs). The deduced amino acid sequence of ORF1, designated ORFtnp, showed high levels of identity with the first 233 amino acids of the putative transposases of the IS1164 elements from Rhodococcus rhodochrous (71%) and Mycobacterium bovis (68%), members of the Staphylococcus aureus IS256 family of transposases. No significant homologies to ORF2 and ORF3 were found in the nucleic acid and protein databases. ORFtnp is located 5' of ORF3. ORF2 is incomplete and is located 3' of ORF3. Subcloning of the individual ORFs demonstrated that ORF3, designated nec1, is sufficient for necrotizing activity in S. lividans 66 TK24. S. lividans 66 TK24 expressing nec1 does not produce thaxtomin A but produces an unidentified extracellular water-soluble compound that causes necrosis on potato tuber discs. The G+C content of nec1 suggests that it has moved horizontally from another genus. Southern analysis of ORFtnp and nec1 demonstrate that these genes are physically linked in Streptomyces strains, including S. scabies and Streptomyces acidiscabies strains, that are pathogenic on potato and that produce the phytotoxin thaxtomin A. These data suggest that nec1 may have been mobilized into S. scabies through a transposition event mediated by ORFtnp. 相似文献
5.
6.
The nucleotide sequence of the 3' end of the nifU coding sequence, the complete coding sequence of nifH and a substantial part of the 5' end of nifD coding sequence from Nostoc 6720 is presented. The coding sequences are highly conserved with those of Anabaena 7120 and Anabaena sp. L31. However the intergenic region between nifU and nifH contains two segments of short tandemly repetitive repeat sequences (STRRs) that differ from the STRR that is common to both Anabaena7120 and Anabaenasp. L31. Various sequence structures that are common to Nostoc 6720, the Anabaena strains and Plectonema boryanum are discussed. 相似文献
7.
A gene (eprA1) encoding the extracellular protease of Aeromonas hydrophila AH1 has been cloned and sequenced. Nucleotide sequence analysis of eprA1 predicted a single open reading frame (ORF) of 1038 bp encoding a 346 amino acid (aa) polypeptide, with a potential 21-aa signal peptide. When the eprA1 gene was expressed in minicells, one major band of approx. 37 kDa was identified, while protease activity staining experiments identified a caseinolytic band of approx. 29 kDa determined by SDS-PAGE analysis of the minicells. The deduced C-terminal aa region (Arg-290 to Gly-313) showed sequence homology to partial C-terminal sequences of other zinc metalloproteases including Penicillium citrinum metalloprotease (PlnC), Aspergillus oryzae metalloprotease (NpII), Aspergillus flavus metalloprotease (MepA), and Aspergillus fumigatus metalloprotease (Mep20), particularly with respect to zinc-binding residues. 相似文献
8.
The complete gene encoding the human N-methyl-D-aspartate receptor subunit NR1 (NMDAR1) has been isolated on a single cosmid clone. The gene is composed of 21 exons distributed over a total length of about 31 kb. More than 24 kb were sequenced. Exons 4, 20 and 21 are identical in their amino-acid sequence to those exons that are subject to alternative splicing in rat, indicating that all eight NMDAR1 isoforms found in rat will also be expressed in the human brain. Computer analysis of the pre-mRNA sequence revealed no secondary structures stable enough to explain alternative splicing. We suggest that cell-specific factors control expression of different isoforms. The promoter region contains two perfect copies of the recognition sequence for the Drosophila even-skipped protein, indicating that the developmentally regulated expression of NMDAR1 is controlled by a homeobox protein. The complete cosmid clone covering NMDAR1 was mapped to chromosome 9q34.3-qter by fluorescent in situ hybridization (FISH). The telomeric location is supported by an imperfect (CA)n repeat homologous to a subtelomeric repeat on chromosome 16p. 相似文献
9.
A cDNA fragment encoding the mitochondrial alternative oxidase, the enzyme responsible for cyanide-insensitive and salicylhydroxamic acid (SHAM)-sensitive respiration, from the citric acid-producing fungus Aspergillus niger WU-2223L was cloned and expressed in Escherichia coli as a host strain. Synthetic primers were designed from the conserved nucleotide sequences of the alternative oxidase genes from higher plants and a yeast. The 210-bp DNA fragment was amplified by PCR with these primers using chromosomal DNA of WU-2223L as a template, and was employed to screen a cDNA library of A. niger. One full-length cDNA clone of 1.2 kb was obtained, and was sequenced to reveal that the clone contained an open reading frame (ORF-AOX1) encoding a polypeptide of 351 amino acids. The predicted amino-acid sequence exhibited 50%, 55%, and 52% homology to the alternative oxidases of Hansenula anomala, Neurospora crassa and Sauromatum guttatum, respectively. In the 5'-terminus region of the ORF-AOX1, a mitochondrial targeting motif was found. The whole open reading frame of ORF-AOX1 was ligated to plasmid pKK223-3 to construct the expression vector pKAOX1. The E. coli transformant harboring pKAOX1 showed cyanide-insensitive and SHAM-sensitive respiration, and expression was increased approximately two-fold by the addition of IPTG. These results indicated that the ORF-AOX1 encodes an alternative oxidase of A. niger. 相似文献
10.
A Romero J Caldeira J Legall I Moura JJ Moura MJ Romao 《Canadian Metallurgical Quarterly》1996,239(1):190-196
The crystal structures of the flavodoxin from Desulfovibrio desulfuricans ATCC 27774 have been determined and refined for both oxidized and semi-reduced forms to final crystallographic R-factors of 17.9% (0.8-0.205-nm resolution) and 19.4% (0.8-0.215-nm resolution) respectively. Native flavodoxin crystals were grown from ammonium sulfate with cell constants a = b = 9.59 nm, c=3.37nm (oxidized crystals) and they belong to space group P3(2)21. Semireduced crystals showed some changes in cell dimensions: a = b = 9.51 nm, c=3.35 nm. The three-dimensional structures are similar to other known flavodoxins and deviations are found essentially in the isoalloxazine ring environment. Conformational changes are observed between both redox states and a flip of the Gly61-Met62 peptide bond occurs upon one-electron reduction of the FMN group. These changes influence the redox potential of the oxidized/semiquinone couple. Modulation of the redox potentials is known to be related to the association constant of the FMN group to the protein. The flavodoxin from D. desulfuricans now studied has a large span between E2 (oxidized --> semiquinone) and E1 (semiquinone --> hydroquinone) redox potentials, both these values being substantially more positive within known flavodoxins. A comparison of their FMN environment was made in both oxidation states in order to correlate functional and structural differences. 相似文献
11.
MF Verhagen AJ Pierik RB Wolbert LF Mallée WG Voorhorst WR Hagen 《Canadian Metallurgical Quarterly》1994,225(1):311-319
A spectroelectrochemical study is described of the sixteen hemes in the high-molecular-mass, monomeric cytochrome c (Hmc) from the periplasmic space of Desulfovibrio vulgaris, strain Hildenborough. One of the hemes has special properties. In the oxidized state at pH 7 it is predominantly high-spin, S = 5/2, with a g perpendicular value of less than 6 indicative of quantum-mechanical mixing with a low-lying (800 cm-1) S = 3/2 state; the balance is probably a low-spin derivative. The high-spin heme has an Em.7.5 value of +61 mV. The fifteen other hemes are low-spin bis-histidine coordinated with Em.7.5 values of approximately -0.20 V. Two of these hemes exhibit very anisotropic EPR spectra with a g1 value of 3.65 characteristic for strained bis-histidine coordination. A previous proposal, namely that methionine is coordinated to one of the hemes [Pollock, W.B.R., Loufti, M. Bruschi, M. Rapp-Giles, B.J., Wall, J. & Voordouw, G. (1991) J. Bacteriol. 173, 220] is disproved using spectroscopic evidence. Contrasting electrochemical data sets from two previous studies [Tan, J. & Cowan, J.A. (1990) Biochemistry 29, 4886; Bruschi, M., Bertrand, P., More, C., Leroy, G., Bonicel, J., Haladjian, J., Chottard, G., Pollock, W.B.R. & Voordouw, G. (1992) Biochemistry 31, 3281] are not consistent with our EPR titration results and are not reproducible. Hmc can be reduced by D. vulgaris Fe-hydrogenase in the presence of molecular hydrogen. 相似文献
12.
The gene encoding the Pyrococcus furiosus hyperthermophilic amylopullulanase (APU) was cloned, sequenced, and expressed in Escherichia coli. The gene encoded a single 827-residue polypeptide with a 26-residue signal peptide. The protein sequence had very low homology (17 to 21% identity) with other APUs and enzymes of the alpha-amylase family. In particular, none of the consensus regions present in the alpha-amylase family could be identified. P. furiosus APU showed similarity to three proteins, including the P. furiosus intracellular alpha-amylase and Dictyoglomus thermophilum alpha-amylase A. The mature protein had a molecular weight of 89,000. The recombinant P. furiosus APU remained folded after denaturation at temperatures of < or = 70 degrees C and showed an apparent molecular weight of 50,000 in sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Denaturating temperatures of above 100 degrees C were required for complete unfolding. The enzyme was extremely thermostable, with an optimal activity at 105 degrees C and pH 5.5. Ca2+ increased the enzyme activity, thermostability, and substrate affinity. The enzyme was highly resistant to chemical denaturing reagents, and its activity increased up to twofold in the presence of surfactants. 相似文献
13.
The Brevibacterium lactofermentum EF-P gene, encoding the elongation factor protein P, was cloned and sequenced. According to DNA sequence analysis of this gene, the B. lactofermentum EF-P protein consists of 187 amino acids with a calculated molecular weight of 20,584. Southern hybridization of an internal fragment of the EF-P gene from B. lactofermentum with chromosomal DNAs from different microorganisms reveals that it is a unique gene product in B. lactofermentum and Corynebacterium glutamicum. The EF-P gene was expressed in E. coli using the T7 expression system and the calculated molecular weight of the expressed protein was 23,000. Disruption experiments using an internal fragment of the EF-P gene or a disrupted EF-P gene in suicide plasmids always failed, suggesting that the gene is needed for cell viability. 相似文献
14.
M Wakayama Y Katsuno S Hayashi Y Miyamoto K Sakai M Moriguchi 《Canadian Metallurgical Quarterly》1995,59(11):2115-2119
The gene encoding the D-aminoacylase of Alcaligenes xylosoxydans subsp. xylosoxydans A-6 (Alcaligenes A-6) was cloned and its complete nucleotide sequence was identified. The D-aminoacylase structural gene consists of 1452 nucleotides and encodes 484 amino acid residues. The molecular weight of D-aminoacylase was calculated to be 51,918. This value agreed well with the apparent molecular weight of 52,000 found for the purified enzyme from Alcaligenes A-6 by sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis (PAGE). The N-terminal amino acid sequence (NH2-SQSDSQPFDLLRAG-) predicted by the nucleotide sequence exactly matched those of the purified D-aminoacylase both from Alcaligenes A-6 and from cloned Escherichia coli (E. coli), with the exception of the removal of the N-terminal methionine processed after translation. The purified recombinant enzyme showed almost the same enzymatic properties as the native enzyme from Alcaligenes A-6. Alcaligenes A-6 D-aminoacylase showed 25-29% homology with L-aminoacylases from Bacillus stearothermophilus, porcine and humans. 相似文献
15.
16.
A HindIII fragment of N. gonorrhoeae MS11 DNA coding for DNA methyltransferase (MTase) activity was cloned and expressed in E. coli AP1-200-9 cells. The sequence of 4681 bp was determined, and its analysis revealed two open reading frames (ORFs) sharing some similarity with known DNA MTases. ORF1 encodes an active N4mC MTase (M.NgoMV). The enzyme modifies only one strand of double stranded DNA and preferentially recognises the sequence GCCHR although it is able to methylate other sites. The exact recognition sequence cannot be precisely defined due to a relaxed specificity. The second ORF shows high homology to 5mC Mtases, but we were unable to demonstrate DNA methylating activity of its product either in vivo or in vitro. 相似文献
17.
Cerebral cortical maps in adult primates reorganize within minutes-hours after peripheral injuries, but subcortical versus intracortical contributions to this rapid reorganization remain controversial. The present results show that injury of nerves to the hands of adult monkeys triggers rapid (minutes-hours) changes in maps of the hand in the brainstem main cuneate nucleus. These findings suggest that peripheral injury causes an initial concurrent reorganization of brainstem and cortical substrates and that early sensory changes emerge from reorganization involving multiple central levels. 相似文献
18.
P Wang CD Denoya MR Morgenstern DD Skinner KK Wallace R Digate S Patton N Banavali G Schuler MK Speedie KA Reynolds 《Canadian Metallurgical Quarterly》1996,178(23):6873-6881
We report the cloning of the gene encoding the 1-cyclohexenylcarbonyl coenzyme A reductase (ChcA) of Streptomyces collinus, an enzyme putatively involved in the final reduction step in the formation of the cyclohexyl moiety of ansatrienin from shikimic acid. The cloned gene, with a proposed designation of chcA, encodes an 843-bp open reading frame which predicts a primary translation product of 280 amino acids and a calculated molecular mass of 29.7 kDa. Highly significant sequence similiarity extending along almost the entire length of the protein was observed with members of the short-chain alcohol dehydrogenase superfamily. The S. collinus chcA gene was overexpressed in Escherichia coli by using a bacteriophage T7 transient expression system, and a protein with a specific ChcA activity was detected. The E. coli-produced ChcA protein was purified and shown to have similar steady-state kinetics and electrophoretic mobility on sodium dodecyl sulfate-polyacrylamide gels as the enoyl-coenzyme A reductase protein prepared from S. collinus. The enzyme demonstrated the ability to catalyze, in vitro, three of the reductive steps involved in the formation of cyclohexanecarboxylic acid. An S. collinus chcA mutant, constructed by deletion of a genomic region comprising the 5' end of chcA, lost the ChcA activity and the ability to synthesize either cyclohexanecarboxylic acid or ansatrienin. These results suggest that chcA encodes the ChcA that is involved in catalyzing multiple reductive steps in the pathway that provides the cyclohexanecarboxylic acid from shikimic acid. 相似文献
19.
Two mutants of the Desulfovibrio vulgaris flavodoxin, T12H and N14H, were generated which, for the first time, place a basic residue within the normally neutral 5'-phosphate binding loop of the flavin mononucleotide cofactor binding site found in all flavodoxins. These histidine residues were designed to form an ion pair with the dianionic 5'-phosphate, either altering its ionization state or offsetting its negative charge to allow evaluation of the magnitude of its electrostatic effect on the redox properties of the cofactor. The midpoint potential for the oxidized/semiquinone couple was not significantly altered in either mutant. However, the midpoint potentials for the semiquinone/hydroquinone couple (Esq/hq) were less negative than that of the wild type, increasing by 28 and 15 mV relative to that of the wild type for the T12H and N14H mutants, respectively, at pH 6. 31P NMR spectroscopy suggests that, just as for wild type, the phosphate group in each mutant does not change its ionization state between pH 6 and 8. Therefore, the small increases in midpoint potential must be linked to the protonation of the histidine residues, either through favorable interactions with the anionic hydroquinone or by the partial compensation of the charge on the 5'-phosphate. Values for the pKa of His12 and His14 in the oxidized flavodoxin were determined by 1H NMR spectroscopy to be 6.71 and 6.93, respectively, which are only modestly elevated relative to the average value for histidines in proteins. This suggests that the histidines do not form strong ion-pairing interactions with the phosphate and/or that the effective charge on the 5'-phosphate may be substantially less than the reported formal dianionic charge. Either way, the data provide evidence for the rather weak electrostatic interaction between a charged group at this site and the anionic flavin hydroquinone. In contrast, Esq/hq reported for the apoflavodoxin-riboflavin complex, which lacks the 5'-phosphate group, is 180 mV less negative than that of the native flavodoxin. The re-evaluation of the redox and cofactor binding properties of the riboflavin complex generated values for the dissociation constants for the riboflavin complex in the oxidized, semiquinone, and hydroquinone oxidation states that are 2100-, 63000-, and 54-fold higher, respectively, than that for the naturally occurring flavin mononucleotide complex. The large redox potential shifts observed for both redox couples in the riboflavin complex are primarily the consequence of a decreased stabilization of the semiquinone rather than the result of the absence of the negative charge of the 5'-phosphate. It is concluded from this study that the negative charge on the phosphate group of the cofactor does not play a disproportionate role in decreasing Esq/hq, at most contributing equivalently with the acidic amino acid residues clustered around the flavin to an unfavorable electrostatic environment for the formation of the flavin hydroquinone anion. 相似文献