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 0 毫秒
1.
2.
3.
The Desulfovibrio gigas aldehyde-oxido-reductase contains molybdenum and iron-sulfur clusters. M?ssbauer spectroscopy was used to characterize the iron-sulfur clusters. Spectra of the enzyme in its oxidized, partially reduced and benzaldehyde-reacted states were recorded at different temperatures and applied magnetic fields. All the iron atoms in D. gigas aldehyde oxido-reductase are organized as [2Fe-2S] clusters. In the oxidized enzyme, the clusters are diamagnetic and exhibit a single quadrupole doublet with parameters (delta EQ = 0.62 +/- 0.02 mm/s and delta = 0.27 +/- 0.01 mm/s) typical for the [2Fe-2S]2+ state. M?ssbauer spectra of the reduced clusters also show the characteristics of a [2Fe-2S]1+ cluster and can be explained by a spin-coupling model proposed for the [2Fe-2S] cluster where a high-spin ferrous ion (S = 2) is antiferromagnetically coupled to a high-spin ferric ion (S = 5/2) to form a S = 1/2 system. Two ferrous sites with different delta EQ values (3.42 mm/s and 2.93 mm/s at 85 K) are observed for the reduced enzyme, indicating the presence of two types of [2Fe-2S] clusters in the D. gigas enzyme. Taking this observation together with the re-evaluated value of iron content (3.5 +/- 0.1 Fe/molecule), it is concluded that, similar to other Mo-hydroxylases, the D. gigas aldehyde oxido-reductase also contains two spectroscopically distinguishable [2Fe-2S] clusters.  相似文献   

4.
Adenylylsulfate reductase (adenosine 5′-phosphosulfate [APS] reductase [APSR]) plays a key role in catalyzing APS to sulfite in dissimilatory sulfate reduction. Here, we report the crystal structure of APSR from Desulfovibrio gigas at 3.1-Å resolution. Different from the α2β2-heterotetramer of the Archaeoglobus fulgidus, the overall structure of APSR from D. gigas comprises six αβ-heterodimers that form a hexameric structure. The flavin adenine dinucleotide is noncovalently attached to the α-subunit, and two [4Fe-4S] clusters are enveloped by cluster-binding motifs. The substrate-binding channel in D. gigas is wider than that in A. fulgidus because of shifts in the loop (amino acid 326 to 332) and the α-helix (amino acid 289 to 299) in the α-subunit. The positively charged residue Arg160 in the structure of D. gigas likely replaces the role of Arg83 in that of A. fulgidus for the recognition of substrates. The C-terminal segment of the β-subunit wraps around the α-subunit to form a functional unit, with the C-terminal loop inserted into the active-site channel of the α-subunit from another αβ-heterodimer. Electrostatic interactions between the substrate-binding residue Arg282 in the α-subunit and Asp159 in the C terminus of the β-subunit affect the binding of the substrate. Alignment of APSR sequences from D. gigas and A. fulgidus shows the largest differences toward the C termini of the β-subunits, and structural comparison reveals notable differences at the C termini, activity sites, and other regions. The disulfide comprising Cys156 to Cys162 stabilizes the C-terminal loop of the β-subunit and is crucial for oligomerization. Dynamic light scattering and ultracentrifugation measurements reveal multiple forms of APSR upon the addition of AMP, indicating that AMP binding dissociates the inactive hexamer into functional dimers, presumably by switching the C terminus of the β-subunit away from the active site. The crystal structure of APSR, together with its oligomerization properties, suggests that APSR from sulfate-reducing bacteria might self-regulate its activity through the C terminus of the β-subunit.Sulfate-reducing bacteria (SRB) are a special group of prokaryotes that are found in sulfate-rich environments because of their ability to metabolize sulfate. SRB use sulfate as the final electron acceptor in various anaerobic environments, such as soil, oil fields, the sea, or the innards of animals or even human beings (10, 11, 19, 25, 33). Their ability to degrade sulfate offers protection against environmental pollution. SRB can remove sulfate and toxic heavy atoms from factory waste waters (12). The Desulfovibrio species is a much-studied representative of SRB, and Desulfovibrio gigas has been studied under many diverse conditions to elucidate metabolic pathways (23, 35).Sulfate reduction is one of the oldest forms of cellular metabolism. The reduction can be either assimilatory or dissimilatory. Sulfate is the terminal electron acceptor in dissimilatory reduction and the raw material for the biosynthesis of cysteine in assimilatory reduction. The latter type of reduction occurs in archaebacteria, bacteria, fungi, and plants via various pathways (17). For example, in Escherichia coli, the reduction initially catalyzes sulfate to adenosine 5′-phosphosulfate (APS) by ATP sulfurylase. APS is then phosphorylated by APS kinase to 3′-phosphate APS, which is then further reduced to sulfite by 3′-phosphate APS reductase (APSR). Finally, sulfite is reduced by sulfite reductase to sulfide, which condenses with O-acetylserine by O-acetylserine lyase to form cysteine. For comparison, in dissimilatory sulfate reduction, sulfate is first catalyzed by ATP sulfurylase to APS, which is then directly reduced by APSR to sulfite. Sulfite is subsequently reduced by dissimilatory sulfite reductase to the following three possible products: trithionite (S3O62−), thiosulfate (S2O32−), or sulfide (S2−).Adenylylsulfate reductase, also called APSR, plays an important role in catalyzing APS to AMP and sulfite in the dissimilatory sulfate reduction. APSR was first partially purified and characterized from Desulfovibrio desulfuricans (32). Multiple forms of APSR in Desulfovibrio vulgaris were observed in buffers under varied conditions (1) and were found in the cytoplasm of cells (18). APSR from D. gigas was first purified by Lampreia et al. (21) and showed a molecular mass of 400 kDa comprised of α- and β-subunits, corresponding to the molecular masses of 70 kDa and 23 kDa, respectively. One flavin adenine dinucleotide (FAD) and two [4Fe-4S] clusters per APSR have been observed and characterized by electron paramagnetic resonance and Mössbauer spectroscopy. The enzyme from D. gigas has been described as an α2β complex involving one FAD and two [4Fe-4S] clusters (20). In D. vulgaris, APSR is apparently an α2β2 complex with a molecular mass of 186 kDa; only one Fe-S cluster is found in the αβ-heterodimer (31). Thus, the subunit and quaternary structures of APSR and their constitution of cofactors in terms of FAD and iron-sulfur clusters are still under debate. Only the enzyme from Archaeoglobus fulgidus has benefited from having an X-ray crystal structure. In this APSR, the functional unit has been shown to be the 1:1 αβ-heterodimer, containing two iron-sulfur clusters and one FAD in the structure (7). However, crystal packing shows that the asymmetric unit is an α2β2-heterotetramer.The catalytic mechanism of APSR can be divided into the transport of electrons and the cleavage of APS by FAD. Electron input to the FAD catalyzes the cleavage of APS, releasing AMP and sulfite. Although there have been a number of mechanisms proposed to explain the catalytic cleavage of APS to AMP and sulfite (7, 8, 13, 20, 34), many features of the postulated mechanism remain unsettled, including the proteinogenic hydrogen acceptor in the reaction, the conformational change in the enzyme induced by reduction/oxidation of the FAD cofactor, and the reasons for the observed multiple forms of APSR. The divergence between A. fulgidus and Desulfovibrio species also suggests an obvious distinction in the phylogeny of the α- and β-subunits of APSR.To clarify the difference between APSR from A. fulgidus and that from Desulfovibrio species, we have undertaken a structural study of APSR from D. gigas for comparison with the A. fulgidus enzyme. We have isolated and purified APSR directly from massive, anaerobically grown D. gigas cells for structure determination and characterization. The comparison of the structures and sequences revealing the notable differences at the C termini, activity sites, and other regions for the function is discussed. The structure of oxidized APSR from D. gigas provides much direct evidence about the subunit interactions and the role of the quaternary structure in the regulation of the catalytic mechanism.  相似文献   

5.
The structures of the hybrid cluster proteins (HCPs) from the sulfate-reducing bacteria Desulfovibrio desulfuricans (ATCC 27774) and Desulfovibrio vulgaris (Hildenborough) have been elucidated at a resolution of 1.25 A using X-ray synchrotron radiation techniques. In the case of the D. desulfuricans protein, protein isolation, purification, crystallization and X-ray data collection were carried out under strict anaerobic conditions, whereas for the D. vulgaris protein the conditions were aerobic. However, both structures are essentially the same, comprising three domains and two iron-sulfur centres. One of these centres situated near the exterior of the molecules in domain 1 is a cubane [4Fe-4S] cluster, whereas the other, located at the interface of the three domains, contains the unusual four-iron cluster initially found in the D. vulgaris protein. Details of the structures and the associated EPR spectroscopy of the D. desulfuricans protein are reported herein. These structures show that the nature of the hybrid cluster, containing both oxygen and sulfur bridges, is independent of the presence of oxygen in the isolation and crystallization procedure and also does not vary significantly with changes in the oxidation state. The structures and amino acid sequences of the HCP are compared with the recently elucidated structure of the catalytic subunit of a carbon monoxide dehydrogenase from Carboxydothermus hydrogenoformans and related dehydrogenases. Electronic supplementary material to this paper can be obtained by using the Springer Link server located at http://dx.doi.org/10.1007/s00775-001-0326-y.  相似文献   

6.
7.
The structure of the sulphydryl protease, actinidin, after refinement at 1.7 Å resolution, is described. The positions of most of the 1666 atoms have been determined with an accuracy better than 0.1 Å; only two residues (219 and 220) and the side-chain of a third (87) cannot be seen. In addition, the model contains 272 solvent molecules, all taken as water, except one which may be an ammonium ion. Atomic B values give a good indication of the mobility of different parts of the structure. Actinidin has a double domain structure, with one domain mostly helical in its secondary structure, and the other domain built around a twisted β-sheet. The geometry of hydrogen bonds in helices, β-structure and turns has been analysed. All are significantly non-linear, with the angle N-?…O ~160 °. Carbonyl groups are tilted outwards from the axis of each helix, the tilting apparently unaffected by whether or not additional hydrogen bonds are made (e.g. to water or side-chain atoms). Each domain is folded round a substantial core of non-polar side-chains, but the interface between domains is mostly polar. Interactions across this interface involve a network of eight buried water molecules, the buried carboxyl and amino groups of Glu35, Glu50, Lys181 and Lys17, other polar side-chains and a few hydrophobic groups. One other internal charged side-chain, that of Glu52, is adjacent to a buried solvent molecule, probably an ammonium ion. Other side-chain environments are described. One proline residue has a cis configuration. The sulphydryl group is oxidized, probably to SO2?, with one oxygen atom clearly visible but the other somewhat less certain. The active site geometry is otherwise compatible with the mechanism proposed by Drenth et al. (1975,1976) for papain. The positions of the 272 solvent molecules are described. The best-ordered water molecules are those that are internal (total of 17), in surface pockets, or in the intermolecular contact regions. These generally form three or four hydrogen bonds, two to proton acceptors and one or two to proton donors. Other water molecules make water bridges on the surface, sometimes covering the exposed edges of non-polar groups. Intermolecular contacts involve few protein atoms, but many water molecules.  相似文献   

8.
The primary and three-dimensional structures of a [NiFe] hydrogenase isolated from D. desulfitricans ATCC 27774 were determined, by nucleotide analysis and single-crystal X-ray crystallography. The three-dimensional structural model was refined to R=0.167 and Rfree=0.223 using data to 1.8 A resolution. Two unique structural features are observed: the [4Fe-4S] cluster nearest the [NiFe] centre has been modified [4Fe-3S-3O] by loss of one sulfur atom and inclusion of three oxygen atoms; a three-fold disorder was observed for Cys536 which binds to the nickel atom in the [NiFe] centre. Also, the bridging sulfur atom that caps the active site was found to have partial occupancy, thus corresponding to a partly activated enzyme. These structural features may have biological relevance. In particular, the two less-populated rotamers of Cys536 may be involved in the activation process of the enzyme, as well as in the catalytic cycle. Molecular modelling studies were carried out on the interaction between this [NiFe] hydrogenase and its physiological partner, the tetrahaem cytochrome c3 from the same organism. The lowest energy docking solutions were found to correspond to an interaction between the haem IV region in tetrahaem cytochrome c3 with the distal [4Fe-4S] cluster in [NiFe] hydrogenase. This interaction should correspond to efficient electron transfer and be physiologically relevant, given the proximity of the two redox centres and the fact that electron transfer decay coupling calculations show high coupling values and a short electron transfer pathway. On the other hand, other docking solutions have been found that, despite showing low electron transfer efficiency, may give clues on possible proton transfer mechanisms between the two molecules.  相似文献   

9.
The 1.2?Å resolution crystal structure of the 29?kDa di-tetrahaem cytochrome c 3 from the sulfate reducing bacterium Desulfovibrio gigas was solved by ab initio methods, making this the largest molecule to be solved by this procedure. The actual refined model of the cysteine-linked dimeric molecule reveals that this molecule is very similar to the non-covalently linked symmetrical dimer of the di-tetrahaem cytochrome c 3 from Desulfomicrobium norvegicum. Each monomer has the typical polypeptide fold, haem arrangement and iron coordination found for the tetrahaem cytochrome c 3 molecules. The interface between the covalently linked monomers in the asymmetric unit of the crystal shows a pseudo two-fold arrangement, disturbed from symmetry by crystal packing forces. The fact that D. gigas contains a dimeric tetrahaem cytochrome with solvent accessible disulfide bridges and that this cytochrome specifically couples hydrogen oxidation to thiosulfate reduction in bacterial extracts provides an interesting aspect related to disulfide exchange reactions in this microorganism.  相似文献   

10.
11.
We report the 98% assignment of the apo-form of an orange protein, containing a novel Mo–Cu cluster isolated from Desulfovibrio gigas. This protein presents a region where backbone amide protons exchange fast with bulk solvent becoming undetectable. These residues were assigned using 13C-detection experiments.  相似文献   

12.
The structure of the protein subunit of satellite tobacco necrosis virus has been solved at 3.7 Å resolution. We have now crystallographically refined the original model and extended the resolution to 2.5 Å in order to get a model accurate enough to explain the details of the subunit interactions. The refinement was done with a novel method utilizing the icosahedral symmetry of the virus particle.The final model shows a complicated network of interactions, involving salt linkages, hydrogen bonds and hydrophobic contacts. In addition, we have located three different metal ion sites in the protein shell, linking the protein subunits together. These sites are probably occupied by calcium ions. One site is found in a general position near the icosahedral 3-fold axis of the virus. The ligands form an octahedral arrangement, with two main chain carbonyl oxygens (O-61 and O-64), one carboxylate oxygen (OD1 from Asp194) of the same subunit and a second carboxylate oxygen (OE1 of Glu25) from a 3-fold related subunit. Two water molecules complete the octahedral arrangement. A second site is on the icosahedral 3-fold axis and is liganded by the carboxylate oxygens of the 3-fold related Asp55 residues. The third metal ion site is found on the 5-fold axis, liganded by the five carbonyl oxygens of Thr138 and two water molecules.We are unable to locate the first 11 N-terminal amino acid residues, which point into the virus interior. No interpretable density for RNA has been found, indicating that the nucleic acid of the virus does not have a unique orientation in the crystal.  相似文献   

13.
Highlights? Improved models by DEN refinement at 7 Å ? Best method is DEN refinement with initial segmented rigid-body refinement ? Rfree has predictive power at 7 Å  相似文献   

14.
Highlights? The 4.5 Å cryo-EM structure of AAV-DJ fully resolves the polypeptide backbone ? Liver tropism selected for in AAV-DJ has not changed the heparin binding site ? Changed conformation in an antigenic loop blocks binding of a neutralizing mAb ? Changed in vivo tropism may result from changed immune interactions  相似文献   

15.
16.
A monomeric nine-haem cytochrome c (9Hcc) with 292 amino acid residues was isolated from cells of the sulfate- and nitrate-reducing bacterium Desulfovibrio desulfuricans ATCC 27774 grown under both nitrate- and sulfate-respiring conditions. The nucleotide sequence encoding the 292 residues was determined, allowing the correction of about 10% of the previous primary structure, determined from 1.8?Å electron density maps. The refinement at 1.8?Å resolution of the structural model was completed, giving an R-value of 16.5%. The nine haem groups are arranged into two tetrahaem clusters, located at both ends of the molecule, with Fe-Fe distances and local protein fold very similar to tetrahaem cytochromes c 3, and the extra haem is located asymmetrically between the two regions. The new primary sequence determination confirmed the 39% sequence homology found between this cytochrome and the C-terminal region (residues 229–514) of the high-molecular-weight cytochrome c (Hmc) from D. vulgaris Hildenborough, providing strong evidence of structural similarity between 9Hcc and the C-terminal region of Hmc. The interaction between 9Hcc and the tetrahaem cytochrome c 3 from the same organism was studied by modelling methods, and the results suggest that a specific interaction is possible between haem 4 of tetrahaem cytochrome c 3 and haem 1 or haem 2 of 9Hcc, in agreement with previous kinetic experiments which showed the catalytic effect of the tetrahaem cytochrome c 3 upon the reduction of 9Hcc by the [NiFe] hydrogenase from D. desulfuricans ATCC 27774. These studies suggest a role for 9Hcc as part of the assembly of redox proteins involved in recycling the molecular hydrogen released by the cell as a result of substrate oxidation.  相似文献   

17.
We present here a comparison of three different X-ray crystal structures of DNA tetradecamer sequence d(CCCGGGTACCCGGG)2 all at about 1.7 Å resolution. The sequence was designed as an attempt to form a DNA four-way junction with A-type helical arms. However, in the presence of zinc, magnesium, and in the absence of any metal ion, it does not take up the junction structure, but forms an A-type double helix. This allowed us to study possible conformational changes in the double helix due to the presence of metal ions. Upon addition of the zinc ion, there is a change in the space group from P41212 to P41. The overall conformation of the duplex remains the same. There are small changes in the interaction of the metal ions with the DNA. In the zinc-bound structure, there are two zinc ions that show direct interaction with the N7 atoms of terminal G13 bases at either end of the molecule. There are small changes in the interhelical contacts. The consequence of these differences is to break some of the symmetry and change the space group.  相似文献   

18.
Amoebocyte is the single type of cell circulating in the horseshoe crab hemolymph, which plays a major role in the defense system of the animal. Granules present in these cells are sensitive to nanogram quantities of bacterial endotoxins, which form the basis of the Limulus amoebocyte lysate (LAL) test. Normally, amoebocytes for the production of the LAL are collected by cardiac puncture; hence, development of the in vitro culture system for amoebocytes will reduce the variability of the lysate and help to conserve the 400 million-yr-old living fossil. In the present investigation we have attempted organ culture of gill flaps that have been shown to be the source of amoebocytes. The gill flaps were cultured at 28 degrees C on a rocker platform in a modified L-15 medium supplemented with 10% v/v horseshoe crab serum. This led to the release of amoebocytes outside the gill flaps for a period of 6-8 wk with a more or less steady number of amoebocytes during the weekly harvest. No significant difference was seen in the yield of amoebocytes from male and female horseshoe crabs. Confocal laser microscopy studies revealed significant difference in the size of amoebocytes released in vitro as compared with those obtained in vivo. Thus, we have optimized the culture conditions for the long-term generation of amoebocytes in vitro from the Indian horseshoe crab Tachypleus gigas by reducing the incidence of contamination, simulating in vivo conditions for the organ culture of gill flaps, and improvising the nutritional status using the modified L-15 medium, providing the desired osmolarity and pH.  相似文献   

19.
Here we present the structure of the T1 domain derived from the voltage-dependent potassium channel Kv1.3 of Homo sapiens sapiens at 1.2 Å resolution crystallized under near-physiological conditions. The crystals were grown without precipitant in 150 mM KPi, pH 6.25. The crystals show I4 symmetry typical of the natural occurring tetrameric assembly of the single subunits. The obtained structural model is based on the highest resolution currently achieved for tetramerization domains of voltage-gated potassium channels. We identified an identical fold of the monomer but inside the tetramer the single monomers show a significant rotation which leads to a different orientation of the tetramer compared to other known structures. Such a rotational movement inside the tetrameric assembly might influence the gating properties of the channel. In addition we see two distinct side chain configurations for amino acids located in the top layer proximal to the membrane (Tyr109, Arg116, Ser129, Glu140, Met142, Arg146), and amino acids in the bottom layer of the T1-domain distal from the membrane (Val55, Ile56, Leu77, Arg86). The relative populations of these two states are ranging from 50:50 for Val55, Tyr109, Arg116, Ser129, Glu140, 60:40 for Met142, 65:35 for Arg86, 70:30 for Arg146, and 80:20 for Ile56 and Leu77. The data suggest that in solution these amino acids are involved in an equilibrium of conformational states that may be coupled to the functional states of the whole potassium channel.  相似文献   

20.
Members of the Psb28 family of proteins are accessory factors implicated in the assembly and repair of the photosystem II complex. We present here the crystal structure of the Psb28 protein (Tlr0493) found in the thermophilic cyanobacterium Thermosynechococcus elongatus at a resolution of 2.3 Å. Overall the crystal structure of the Psb28 monomer is similar to the solution structures of C-terminally His-tagged Psb28-1 from Synechocystis sp. PCC 6803 obtained previously by nuclear magnetic resonance spectroscopy. One new aspect is that Escherichia coli-expressed T. elongatus Psb28 is able to form dimers in solution and packs as a dimer of dimers in the crystal. Analysis of wild type and mutant strains of Synechocystis 6803 by blue native-polyacrylamide gel electrophoresis suggests that Psb28-1, the closest homologue to T. elongatus Psb28 in this organism, also exists as an oligomer in vivo, most likely a dimer. In line with the prediction based on the crystal structure of T. elongatus Psb28, the addition of a 3× Flag-tag to the C-terminus of Synechocystis 6803 Psb28-1 interferes with the accumulation of the Psb28-1 oligomer in vivo. In contrast, the more distantly related Psb28-2 protein found in Synechocystis 6803 lacks the residues that stabilize dimer formation in the T. elongatus Psb28 crystal and is detected as a monomer in vivo. Overall our data suggest that the dimer interface in the Psb28 crystal might be physiologically relevant.  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