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 46 毫秒
1.
The diheme cytochrome NapB constitutes the small subunit of a periplasmic nitrate reductase found in a wide variety of bacterial species, including pathogens. The NapB protein is essential in transferring electrons to the large catalytic subunit NapA, which subsequently reduces nitrate to nitrite. Here we present the crystal structure of a proteolyzed form of recombinant NapB from Haemophilus influenzae, which was determined by the multiple-wavelength anomalous dispersion (MAD) method at 1.25 A resolution. This structure shows an unprecedented fold, confirming that NapB proteins belong to a new class of cytochromes. The two heme groups have nearly parallel heme planes and are stacked at van der Waals distances with an iron-to-iron distance of only 9.9 A, two structural features that are also present in the split-Soret diheme cytochrome c from Desulfovibrio desulfuricans ATCC 27774, which is otherwise unrelated in the peptide chain folding pattern. The two propionate side chains on both heme groups are hydrogen-bonded to each other, a structural characteristic that to date also has not been reported in any other heme protein. The propionates of one of the heme groups are pulled toward the interior of the molecule due to a salt bridge and a number of hydrogen bonds between the propionates and conserved residues. We propose a hypothetical but plausible model of the NapAB complex in which the four redox centers are positioned in a virtually linear configuration which spans a distance of nearly 40 A, suggesting an efficient pathway for the transfer of electrons from NapC, the physiological electron donor of NapB, to a nitrate molecule at the catalytic site of NapA.  相似文献   

2.
The structure of the respiratory nitrate reductase (NapAB) from Rhodobacter sphaeroides, the periplasmic heterodimeric enzyme responsible for the first step in the denitrification process, has been determined at a resolution of 3.2 A. The di-heme electron transfer small subunit NapB binds to the large subunit with heme II in close proximity to the [4Fe-4S] cluster of NapA. A total of 57 residues at the N- and C-terminal extremities of NapB adopt an extended conformation, embracing the NapA subunit and largely contributing to the total area of 5,900 A(2) buried in the complex. Complex formation was studied further by measuring the variation of the redox potentials of all the cofactors upon binding. The marked effects observed are interpreted in light of the three-dimensional structure and depict a plasticity that contributes to an efficient electron transfer in the complex from the heme I of NapB to the molybdenum catalytic site of NapA.  相似文献   

3.
The Escherichia coli NapA (periplasmic nitrate reductase) contains a [4Fe-4S] cluster and a Mo-bis-molybdopterin guanine dinucleotide cofactor. The NapA holoenzyme associates with a di-heme c-type cytochrome redox partner (NapB). These proteins have been purified and studied by spectropotentiometry, and the structure of NapA has been determined. In contrast to the well characterized heterodimeric NapAB systems ofalpha-proteobacteria, such as Rhodobacter sphaeroides and Paracoccus pantotrophus, the gamma-proteobacterial E. coli NapA and NapB proteins purify independently and not as a tight heterodimeric complex. This relatively weak interaction is reflected in dissociation constants of 15 and 32 mum determined for oxidized and reduced NapAB complexes, respectively. The surface electrostatic potential of E. coli NapA in the apparent NapB binding region is markedly less polar and anionic than that of the alpha-proteobacterial NapA, which may underlie the weaker binding of NapB. The molybdenum ion coordination sphere of E. coli NapA includes two molybdopterin guanine dinucleotide dithiolenes, a protein-derived cysteinyl ligand and an oxygen atom. The Mo-O bond length is 2.6 A, which is indicative of a water ligand. The potential range over which the Mo(6+) state is reduced to the Mo(5+) state in either NapA (between +100 and -100 mV) or the NapAB complex (-150 to -350 mV) is much lower than that reported for R. sphaeroides NapA (midpoint potential Mo(6+/5+) > +350 mV), and the form of the Mo(5+) EPR signal is quite distinct. In E. coli NapA or NapAB, the Mo(5+) state could not be further reduced to Mo(4+). We then propose a catalytic cycle for E. coli NapA in which nitrate binds to the Mo(5+) ion and where a stable des-oxo Mo(6+) species may participate.  相似文献   

4.
In eukaryotic cells, importin alpha is the major carrier for transport protein cargoes into the nucleus. We characterize here kapA, the single Aspergillus nidulans gene encoding an importin alpha. Using an affinity approach, we identify six potential interactors of KapA(50), a deleted version of KapA lacking the autoinhibitory importin-beta-binding domain. One such interactor is NapB, the A. nidulans orthologue of Saccharomyces cerevisiae Vps75p, a histone chaperone member of the Nap/SET family of proteins that additionally plays a cytosolic role in vacuolar protein sorting. NapB, but not its close relative NapA (the A. nidulans orthologue of yeast Nap1p) interacts directly with KapA(50) in pull down assays, despite the fact that NapB does not contain a classical nuclear localization sequence. NapB is a nuclear protein which exits nuclei at the onset of mitosis when two simultaneous mechanisms might be acting, the partial disassembly of the nuclear pore complexes and as yet unidentified posttranslational modification of NapB. The mitotic cytosolic localization of NapB might facilitate its putative role in the sorting of protein cargoes to the vacuole. In addition, we show that NapB and the mitotic B-type cyclin NimE compete for in vitro binding to KapA.  相似文献   

5.
In eukaryotic cells, importin alpha is the major carrier for transport protein cargoes into the nucleus. We characterize here kapA, the single Aspergillus nidulans gene encoding an importin alpha. Using an affinity approach, we identify six potential interactors of KapA(50), a deleted version of KapA lacking the autoinhibitory importin-beta-binding domain. One such interactor is NapB, the A. nidulans orthologue of Saccharomyces cerevisiae Vps75p, a histone chaperone member of the Nap/SET family of proteins that additionally plays a cytosolic role in vacuolar protein sorting. NapB, but not its close relative NapA (the A. nidulans orthologue of yeast Nap1p) interacts directly with KapA(50) in pull down assays, despite the fact that NapB does not contain a classical nuclear localization sequence. NapB is a nuclear protein which exits nuclei at the onset of mitosis when two simultaneous mechanisms might be acting, the partial disassembly of the nuclear pore complexes and as yet unidentified posttranslational modification of NapB. The mitotic cytosolic localization of NapB might facilitate its putative role in the sorting of protein cargoes to the vacuole. In addition, we show that NapB and the mitotic B-type cyclin NimE compete for in vitro binding to KapA.  相似文献   

6.
The periplasmic nitrate reductase (NAP) from Paracoccus pantotrophus is a soluble two-subunit enzyme (NapAB) that binds two c-type haems, a [4Fe-4S] cluster and a bis-molybdopterin guanine dinucleotide cofactor that catalyses the reduction of nitrate to nitrite. In the present work the NapAB complex has been studied by magneto-optical spectroscopy to probe co-ordination of both the NapB haems and the NapA active site Mo. The absorption spectrum of the NapAB complex is dominated by features from the NapB c-type cytochromes. Using a combination of electron paramagnetic resonance spectroscopy and magnetic circular dichroism it was demonstrated that both haems are low-spin with bis-histidine axial ligation. In addition, a window between 600 and 800 nm was identified in which weak absorption features that may arise from Mo could be detected. The low-temperature MCD spectrum shows oppositely signed bands in this region (peak 648 nm, trough 714 nm) which have been assigned to S-to-Mo(V) charge transfer transitions.  相似文献   

7.
Many sulphate reducing bacteria can also reduce nitrite, but relatively few isolates are known to reduce nitrate. Although nitrate reductase genes are absent from Desulfovibrio vulgaris strain Hildenborough, for which the complete genome sequence has been reported, a single subunit periplasmic nitrate reductase, NapA, was purified from Desulfovibrio desulfuricans strain 27774, and the structural gene was cloned and sequenced. Chromosome walking methods have now been used to determine the complete sequence of the nap gene cluster from this organism. The data confirm the absence of a napB homologue, but reveal a novel six-gene organisation, napC-napM-napA-napD-napG-napH. The NapC polypeptide is more similar to the NrfH subgroup of tetraheme cytochromes than to NapC from other bacteria. NapM is predicted to be a tetra-heme c-type cytochrome with similarity to the small tetraheme cytochromes from Shewanella oneidensis. The operon is located close to a gene encoding a lysyl-tRNA synthetase that is also found in D. vulgaris. We suggest that electrons might be transferred to NapA either from menaquinol via NapC, or from other electron donors such as formate or hydrogen via the small tetraheme cytochrome, NapM. We also suggest that, despite the absence of a twin-arginine targeting sequence, NapG might be located in the periplasm where it would provide an alternative direct electron donor to NapA.  相似文献   

8.
Seven genes, napKEFDABC, encoding the periplasmic nitrate reductase system were cloned from the denitrifying phototrophic bacterium Rhodobacter sphaeroides f. sp. denitrificans IL106. Two transmembrane proteins, NapK and NapE, an iron-sulfur protein NapF, a soluble protein NapD, a catalytic subunit of nitrate reductase precursor NapA, a soluble c-type diheme cytochrome precursor NapB, and a membrane-anchored c-type tetraheme cytochrome NapC were deduced as the gene products. Every mutant in which each nap gene was disrupted by omega-cassette insertion lost nitrate reductase activity as well as the ability of cells to grow with nitrate under anaerobic-dark conditions. A transconjugant of the napD-disrupted mutant with a plasmid bearing the napKEFDABC genes recovered both nitrate reductase activity and nitrate-dependent anaerobic-dark growth of cells. Denitrification activity, which was not observed in the napD mutant, was also restored by the conjugation. These results indicate that the periplasmic nitrate reductase encoded by the napKEFDABC genes is the enzyme responsible for denitrification in this phototroph, although the presence of a membrane-bound nitrate reductase has been reported in the same strain.  相似文献   

9.
We purified the nitrate reductase from the soluble fraction of Magnetospirillum magnetotacticum MS-1. The enzyme was composed of 86- and 17-kDa subunits and contained molybdenum, non-heme iron, and heme c. These properties are very similar to those of the periplasmic nitrate reductase found in Paracoccus pantotrophus. The M. magnetotacticum nap locus was clustered in seven open reading frames, napFDAGHBC. The phylogenetic analyses of NapA, NapB, and NapC suggested a close relationship between M. magnetotacticum nap genes and Escherichia coli nap genes, which is not consistent with the 16S rDNA data. This is the first finding that the alpha subclass of Proteobacteria possesses a napFDAGHBC-type nap gene cluster. The nap gene cluster had putative fumarate and nitrate reduction regulatory protein (Fnr) and NarL protein binding sites. Furthermore, we investigated the effect of molybdate deficiency in medium on the total iron content of the magnetosome fraction and discussed the physiological function of nitrate reductase in relation to the magnetite synthesis in M. magnetotacticum.  相似文献   

10.
It is generally accepted that most gastrointestinal diseases are probably caused by the bacterial pathogen Helicobacter pylori (H. pylori). In this study we have focused on the comparison of protein expression profiles of H. pylori grown under normal and high-salt conditions by a proteomics approach. We have identified about 190 proteins whose expression levels changed after growth at high salt concentration. Among these proteins, neutrophil-activating protein (NapA) was found to be consistently up-regulated under osmotic stress brought by high salts. We have investigated the effect of high salt on secondary and tertiary structures of NapA by circular dichroism spectroscopy followed by analytical ultracentrifugation to monitor the change of quaternary structure of recombinant NapA with increasing salt concentration. The loss of iron-binding activity of NapA coupled with noticeable energetic variation in protein association of NapA as revealed by isothermal titration calorimetry was found under high salt condition. The phylogenetic tree analysis based on sequence comparison of 16 protein sequences encompassing NapA proteins and ferritin of H. pylori and other prokaryotic organisms pointed to the fact that all H. pylori NapA proteins of human origin are more homologous to NapA of Helicobacter genus than to other bacterial NapA. Based on computer modeling, NapA proteins from H. pylori of human isolates are found more similar to ferritin from H. pylori than to NapA from other species of bacteria. Taken together, these results suggested that divergent evolution of NapA and ferritin possessing dissimilar and diverse sequences follows a path distinct from that of convergent evolution of NapA and ferritin with similar dual functionality of iron-binding and ferroxidase activities.  相似文献   

11.
The twin-arginine translocation (Tat) pathway is a protein targeting system present in many prokaryotes. The physiological role of the Tat pathway is the transmembrane translocation of fully-folded proteins, which are targeted by N-terminal signal peptides bearing conserved SRRxFLK 'twin-arginine' amino acid motifs. In Escherichia coli the majority of Tat targeted proteins bind redox cofactors and it is important that only mature, cofactor-loaded precursors are presented for export. Cellular processes have been unearthed that sequence these events, for example the signal peptide of the periplasmic nitrate reductase (NapA) is bound by a cytoplasmic chaperone (NapD) that is thought to regulate assembly and export of the enzyme. In this work, genetic, biophysical and structural approaches were taken to dissect the interaction between NapD and the NapA signal peptide. A NapD binding epitope was identified towards the N-terminus of the signal peptide, which overlapped significantly with the twin-arginine targeting motif. NMR spectroscopy revealed that the signal peptide adopted a α-helical conformation when bound by NapD, and substitution of single residues within the NapA signal peptide was sufficient to disrupt the interaction. This work provides an increased level of understanding of signal peptide function on the bacterial Tat pathway.  相似文献   

12.
The periplasmic nitrate reductase (NapAB) from Cupriavidus necator is a heterodimeric protein that belongs to the dimethyl sulfoxide reductase family of mononuclear Mo-containing enzymes and catalyzes the reduction of nitrate to nitrite. The protein comprises a large catalytic subunit (NapA, 91 kDa) containing the molybdenum active site plus one [4Fe-4S] cluster, as well as a small subunit (NapB, 17 kDa), which is a diheme c-type cytochrome involved in electron transfer. Crystals of the oxidized form of the enzyme diffracted beyond 1.5 Å at the European Synchrotron Radiation Facility. This is the highest resolution reported to date for a nitrate reductase, providing true atomic details of the protein active center, and this showed further evidence on the molybdenum coordination sphere, corroborating previous data on the related Desulfovibrio desulfuricans NapA. The molybdenum atom is bound to a total of six sulfur atoms, with no oxygen ligands or water molecules in the vicinity. In the present work, we were also able to prepare partially reduced crystals that revealed two alternate conformations of the Mo-coordinating cysteine. This crystal form was obtained by soaking dithionite into crystals grown in the presence of the ionic liquid [C4mim]Cl. In addition, UV-Vis and EPR spectroscopy studies showed that the periplasmic nitrate reductase from C. necator might work at unexpectedly high redox potentials when compared to all periplasmic nitrate reductases studied to date.  相似文献   

13.
The periplasmic nitrate reductase (Nap) is wide-spread in proteobacteria. NapA, the nitrate reductase catalytic subunit, contains a Mo-bisMGD cofactor and one [4Fe-4S] cluster. The nap gene clusters in many bacteria, including Rhodobacter sphaeroides DSM158, contain an napF gene, disruption of which drastically decreases both in vitro and in vivo nitrate reductase activities. In spite its importance in the Nap system, NapF has never been characterized biochemically, and its role remains unknown. The NapF protein has four polycysteine clusters that suggest that it is an iron-sulfur-containing protein. In the present study, a His(6)-tagged NapF protein was overproduced in Escherichia coli and purified anaerobically. The purified NapF protein was used to obtain polyclonal antibodies raised in rabbit, and cellular fractionation of R. sphaeroides followed by immunoprobing with anti-NapF antibodies revealed that the native NapF protein is located in the cytoplasm. This contrasts with the periplasmic location of the mature NapA. However, NapA could not be detected in an isogenic napF(-) strain of R. sphaeroides. The His(6)-tagged NapF protein displayed spectral properties indicative of Fe-S clusters, but these features were rapidly lost, suggesting cluster lability. However, reconstitution of the Fe-S centers into the apo-NapF protein was achieved in the presence of Azotobacter vinelandii cysteine desulfurase (NifS), and this allowed the recovery of nitrate reductase activity in NapA protein that had previously been treated with 2,2'-dipyridyl to remove the [4Fe-4S] cluster. This activity was not recovered in the absence of NapF. Taking into account the cytoplasmic localization of NapF, the presence of labile Fe-S clusters in the protein, the napF(-) strain phenotype, and the NapF-dependent reactivation of the 2,2'-dipyridyl-treated NapA, we propose a role for NapF in assembling the [4Fe-4S] center of the catalytic subunit NapA.  相似文献   

14.
The rumen bacterium Wolinella succinogenes grows by respiratory nitrate ammonification with formate as electron donor. Whereas the enzymology and coupling mechanism of nitrite respiration is well known, nitrate reduction to nitrite has not yet been examined. We report here that intact cells and cell fractions catalyse nitrate and chlorate reduction by reduced viologen dyes with high specific activities. A gene cluster encoding components of a putative periplasmic nitrate reductase system (napA, G, H, B, F, L, D) was sequenced. The napA gene was inactivated by inserting a kanamycin resistance gene cassette. The resulting mutant did not grow by nitrate respiration and did not reduce nitrate during growth by fumarate respiration, in contrast to the wild type. An antigen was detected in wild-type cells using an antiserum raised against the periplasmic nitrate reductase (NapA) from Paracoccus pantotrophus. This antigen was absent in the W. succinogenes napA mutant. It is concluded that the periplasmic nitrate reductase NapA is the only respiratory nitrate reductase in W. succinogenes, although a second nitrate-reducing enzyme is apparently induced in the napA mutant. The nap cluster of W. succinogenes lacks a napC gene whose product is thought to function in quinol oxidation and electron transfer to NapA in other bacteria. The W. succinogenes genome encodes two members of the NapC/NirT family, NrfH and FccC. Characterization of corresponding deletion mutants indicates that neither of these two proteins is required for nitrate respiration. A mutant lacking the genes encoding respiratory nitrite reductase (nrfHA) had wild-type properties with respect to nitrate respiration. A model of the electron transport chain of nitrate respiration is proposed in which one or more of the napF, G, H and L gene products mediate electron transport from menaquinol to the periplasmic NapAB complex. Inspection of the W. succinogenes genome sequence suggests that ammonia formation from nitrate is catalysed exclusively by periplasmic respiratory enzymes.  相似文献   

15.
The reduction of nitrate to nitrite in the bacterial periplasm occurs in the 90 kDa NapA subunit of the periplasmic nitrate reductase (NAP) system. Most Shewanella genomes contain two nap operons: napEDABC and napDAGHB, which is an unusual feature of this genus. Two native, monomeric, 90 kDa nitrate reductase active proteins were resolved by hydrophobic interaction chromatography from aerobic cultures of Shewanella gelidimarina replete with reduced nitrogen compounds. The 90 kDa protein obtained in higher yield was characterized as NapA by electronic absorption and electron paramagnetic resonance spectroscopies and was identified by LC/MS/MS and MALDI-TOF/TOF MS as NapA from the napEDABC-type operon. The other 90 kDa protein, which was unstable and produced in low yields, was posited as NapA from the napDAGHB-type operon. Two napA genes have been sequenced from the napEDABC-type and napDAGHB-type operons of S. gelidimarina. Native NAP from S. putrefaciens was resolved as one NapA monomer and one NapAB heterodimer. Two amino acid substitutions in NapA correlated with the isolation of NAP as a NapA monomer or a NapAB heterodimer. The resolution of native, redox-active NapA isoforms in Shewanella provides new insight into the respiratory versatility of this genus, which has implications in bioremediation and the assembly of microbial fuel cells.  相似文献   

16.
The expression of the yeast gene CPA1, which encodes the small subunit of the arginine pathway carbamoylphosphate synthetase, is repressed by arginine at a translational level. CPA1 mRNA contains a 250-nucleotide-long leader which includes a 25-codon upstream open reading frame (uORF). Oligonucleotide site-directed mutagenesis of this uORF as well as sequencing of constitutive cis-dominant mutations has suggested that the leader peptide product of the CPA1 uORF is an essential negative element for repression of the CPA1 gene by arginine. In this work, a series of deletions affecting the regions 5' and 3' to the uORF in the leader sequence was constructed. The arginine-dependent repression of CPA1 was little affected in these constructions, indicating that these regions are not essential for the regulatory response. This conclusion was further supported by the finding that inserting the mRNA segment encoding the leader peptide sequence of CPA1 in the leader sequence of another gene, namely, GCN4, places this gene under arginine repression. Similarly, the behavior of fusions of the leader sequence of CPA1 with those of ARG4 or GAL10 confirmed that the regions of this leader located upstream and downstream from the uORF are dispensable for the regulation by arginine. Finally, a set of substitution mutations which modify the uORF nucleotide sequence while leaving unchanged the corresponding amino acid sequence was constructed. The mutations did not affect the repression of CPA1 by arginine. The data presented in this paper consequently agree with the conclusion that the leader peptide itself is the main element required for the translational repression of CPA1.  相似文献   

17.
Pseudomonas solanacearum is an important phytopathogen that produces a variety of extracellular enzymes. Previous reports suggested that one of these, a 43-kDa beta-1,4-endoglucanase (EGL), is initially synthesized with a 45-residue leader sequence that is removed during export. Experiments with globomycin presented here also suggest that the primary precursor of EGL (ppEGL) has a 45-residue leader sequence but that only the first 19 residues of the leader sequence are removed by signal peptidase II during initial export across the inner membrane. Further analysis suggested that the resultant 46-kDa intermediate precursor (pEGL) is a transient fatty acylated lipoprotein and is located on the periplasmic side of the inner membrane of P. solanacearum. Although Escherichia coli could synthesize ppEGL, modify it with palmitate, and remove the first 19 residues of the leader sequence during export across the inner membrane, only P. solanacearum could export pEGL across the outer membrane and remove the remaining 26 residues of the leader sequence producing the mature, extracellular EGL. The second step of the export process requires export machinery not present in E. coli. To our knowledge this represents the first example of a leader sequence with two distinct parts, one removed during export across the inner membrane and the other removed during export across the outer membrane.  相似文献   

18.
19.
Killer strains of Kluyveromyces lactis secrete a toxin which presumably is processed during secretion from a larger precursor. Analysis of the sequence of the K. lactis killer toxin gene predicts that the first 16 amino acids at the amino terminus of the protein should represent its leader peptide. We have tested the capability of this leader peptide to direct secretion of a protein fused to it by inserting a synthetic oligonucleotide identical to the sequence of the putative leader peptide into a yeast expression vector. Subsequently, the cDNA coding for the secreted active portion of the human interleukin 1 beta (IL-1 beta) was fused to the leader peptide sequence of the killer toxin. This construction in Saccharomyces cerevisiae is capable of directing synthesis and secretion of correctly processed IL-1 beta into the culture medium.  相似文献   

20.
During the replication of equine arteritis virus (EAV) six subgenomic mRNAs are synthesized. We present evidence that the viral mRNAs form a 3'-coterminal nested set and contain a common leader sequence of 208 nucleotides which is encoded by the 5'-end of the genome. The leader is joined to the bodies of mRNA 5 and 6 at positions defined by the sequence 5' UCAAC 3'. The part of the leader sequence flanking the UCAAC motif is very similar to the 5'-splice site of the Tetrahymena pre-rRNA. A possible internal guide sequence has been identified 43 nucleotides downstream of the leader sequence on the genome. Hybridization analysis shows that all EAV intracellular RNAs contain the leader sequence. These data imply that the viral subgenomic mRNAs are composed of leader and body sequences which are non-contiguous on the genome.  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