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.
4.
5.
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   

6.
7.
Activation of voltage-gated sodium (Na(v)) channels initiates and propagates action potentials in electrically excitable cells. β-Scorpion toxins, including toxin IV from Centruroides suffusus suffusus (CssIV), enhance activation of Na(V) channels. CssIV stabilizes the voltage sensor in domain II in its activated state via a voltage-sensor trapping mechanism. Amino acid residues required for the action of CssIV have been identified in the S1-S2 and S3-S4 extracellular loops of domain II. The extracellular loops of domain III are also involved in toxin action, but individual amino acid residues have not been identified. We used site-directed mutagenesis and voltage clamp recording to investigate amino acid residues of domain III that are involved in CssIV action. In the IIISS2-S6 loop, five substitutions at four positions altered voltage-sensor trapping by CssIV(E15A). Three substitutions (E1438A, D1445A, and D1445Y) markedly decreased voltage-sensor trapping, whereas the other two substitutions (N1436G and L1439A) increased voltage-sensor trapping. These bidirectional effects suggest that residues in IIISS2-S6 make both positive and negative interactions with CssIV. N1436G enhanced voltage-sensor trapping via increased binding affinity to the resting state, whereas L1439A increased voltage-sensor trapping efficacy. Based on these results, a three-dimensional model of the toxin-channel interaction was developed using the Rosetta modeling method. These data provide additional molecular insight into the voltage-sensor trapping mechanism of toxin action and define a three-point interaction site for β-scorpion toxins on Na(V) channels. Binding of α- and β-scorpion toxins to two distinct, pseudo-symmetrically organized receptor sites on Na(V) channels acts synergistically to modify channel gating and paralyze prey.  相似文献   

8.
In vertebrate species, the innate immune system down-regulates protein translation in response to viral infection through the action of the double-stranded RNA (dsRNA)-activated protein kinase (PKR). In some teleost species another protein kinase, Z-DNA-dependent protein kinase (PKZ), plays a similar role but instead of dsRNA binding domains, PKZ has Zα domains. These domains recognize the left-handed conformer of dsDNA and dsRNA known as Z-DNA/Z-RNA. Cyprinid herpesvirus 3 infects common and koi carp, which have PKZ, and encodes the ORF112 protein that itself bears a Zα domain, a putative competitive inhibitor of PKZ. Here we present the crystal structure of ORF112-Zα in complex with an 18-bp CpG DNA repeat, at 1.5 Å. We demonstrate that the bound DNA is in the left-handed conformation and identify key interactions for the specificity of ORF112. Localization of ORF112 protein in stress granules induced in Cyprinid herpesvirus 3-infected fish cells suggests a functional behavior similar to that of Zα domains of the interferon-regulated, nucleic acid surveillance proteins ADAR1 and DAI.  相似文献   

9.
10.
The catalytic monoclonal antibody 9A8 (MA 9A8), antiidiotypic to the antibody AE-2 (MA AE2) produced to the active site of acetyl cholinesterase from human erythrocytes, was subjected to a structure–function study. The specific binding of MA 9A8 to MA AE2 (K 2.26 × 109 M–1) was found by the method of surface plasmon resonance, and the functional activity of MA 9A8 was demonstrated. Unlike acetyl cholinesterase, this antibody specifically reacted with the irreversible phosphonate inhibitors of esterases. A peptide map of MA 9A8 was analyzed by MALDI mass spectrometry. The Ser99 residue of its heavy chain was shown to be within the active site of the catalytic antibody. A computer modeling of the MA 9A8 active site suggested the existence of a catalytic dyad formed by Ser99 and His35. A comparison of the tertiary structures of the MA 9A8 and the 17E8 monoclonal antibody, which also exhibited the esterase activity and was produced to the stable analogue of the reaction transition state, indicated a practically complete coincidence of the structures of their presumed active sites.  相似文献   

11.
12.
Eukaryotic cells commonly use protein kinases in signaling systems that relay information and control a wide range of processes. These enzymes have a fundamentally similar structure, but achieve functional diversity through variable regions that determine how the catalytic core is activated and recruited to phosphorylation targets. “Hippo” pathways are ancient protein kinase signaling systems that control cell proliferation and morphogenesis; the NDR/LATS family protein kinases, which associate with “Mob” coactivator proteins, are central but incompletely understood components of these pathways. Here we describe the crystal structure of budding yeast Cbk1–Mob2, to our knowledge the first of an NDR/LATS kinase–Mob complex. It shows a novel coactivator-organized activation region that may be unique to NDR/LATS kinases, in which a key regulatory motif apparently shifts from an inactive binding mode to an active one upon phosphorylation. We also provide a structural basis for a substrate docking mechanism previously unknown in AGC family kinases, and show that docking interaction provides robustness to Cbk1’s regulation of its two known in vivo substrates. Co-evolution of docking motifs and phosphorylation consensus sites strongly indicates that a protein is an in vivo regulatory target of this hippo pathway, and predicts a new group of high-confidence Cbk1 substrates that function at sites of cytokinesis and cell growth. Moreover, docking peptides arise in unstructured regions of proteins that are probably already kinase substrates, suggesting a broad sequential model for adaptive acquisition of kinase docking in rapidly evolving intrinsically disordered polypeptides.  相似文献   

13.
Bis(monoacylglycero)phosphate (BMP) is one of the most intriguing classes of lipids discovered to date. Its biosynthetic pathway is still not fully known, even though studies point to cardiolipin and PG as sources. BMP has an unusual stereoconfiguration which is responsible for its slow catabolism. It is localized exclusively in the inner membranes of late endosomes, and mostly, lysosomes. BMP is negatively charged at the acidic pH of lysosomes, and these charges are central to its role in the degradation of lipids and membranes in the lysosome. In effect, these negative charges facilitate the adhesion of the soluble positively charged hydrolases and activator proteins, thus allowing them to degrade the lipids at the interface of the inner membranes of the lysosome. In this review, we will summarize the different properties of BMP and showcase its importance in the catabolism of lipids and membranes in the lysosome.  相似文献   

14.
Mammalian glutamate dehydrogenase (GDH) catalyzes the reversible inter-conversion of glutamate to α-ketoglutarate and ammonia, interconnecting carbon skeleton and nitrogen metabolism. In addition, it functions as an energy switch by its ability to fuel the Krebs cycle depending on the energy status of the cell. As GDH lies at the intersection of several metabolic pathways, its activity is tightly regulated by several allosteric compounds that are metabolic intermediates. In contrast to other mammals that have a single GDH-encoding gene, humans and great apes possess two isoforms of GDH (hGDH1 and hGDH2, encoded by the GLUD1 and GLUD2 genes, respectively) with distinct regulation pattern, but remarkable sequence similarity (they differ, in their mature form, in only 15 of their 505 amino-acids). The GLUD2 gene is considered a very young gene, emerging from the GLUD1 gene through retro-position only recently (<23 million years ago). The new hGDH2 iso-enzyme, through random mutations and natural selection, is thought to have conferred an evolutionary advantage that helped its persistence through primate evolution. The properties of the two highly homologous human GDHs have been studied using purified recombinant hGDH1 and hGDH2 proteins obtained by expression of the corresponding cDNAs in Sf21 cells. According to these studies, in contrast to hGDH1 that maintains basal activity at 35–40 % of its maximal, hGDH2 displays low basal activity that is highly responsive to activation by rising levels of ADP and/or l-leucine which can also act synergistically. While hGDH1 is inhibited potently by GTP, hGDH2 shows remarkable GTP resistance. Furthermore, the two iso-enzymes are differentially inhibited by estrogens, polyamines and neuroleptics, and also differ in heat-lability. To elucidate the molecular mechanisms that underlie these different regulation patterns of the two iso-enzymes (and consequently the evolutionary adaptation of hGDH2 to a new functional role), we have performed mutagenesis at sites of difference in their amino acid sequence. Results showed that the low basal activity, heat-lability and estrogen sensitivity of hGDH2 could be, at least partially, ascribed to the Arg443Ser evolutionary change, whereas resistance to GTP inhibition has been attributed to the Gly456Ala change. Other amino acid substitutions studied thus far cannot explain all the remaining functional differences between the two iso-enzymes. Also, the Arg443Ser/Gly456Ala double mutation in hGDH1 approached the properties of wild-type hGDH2, without being identical to it. The insights into the structural mechanism of enzymatic regulation and the implications in cell biology provided by these findings are discussed.  相似文献   

15.
Kinesin-5s are microtubule-dependent motors that drive spindle pole separation during mitosis. We used cryo-electron microscopy to determine the 4.5-Å resolution structure of the motor domain of the fission yeast kinesin-5 Cut7 bound to fission yeast microtubules and explored the topology of the motor–microtubule interface and the susceptibility of the complex to drug binding. Despite their non-canonical architecture and mechanochemistry, Schizosaccharomyces pombe microtubules were stabilized by epothilone at the taxane binding pocket. The overall Cut7 footprint on the S. pombe microtubule surface is altered compared to mammalian tubulin microtubules because of their different polymer architectures. However, the core motor–microtubule interaction is tightly conserved, reflected in similar Cut7 ATPase activities on each microtubule type. AMPPNP-bound Cut7 adopts a kinesin-conserved ATP-like conformation including cover neck bundle formation. However, the Cut7 ATPase is not blocked by a mammalian-specific kinesin-5 inhibitor, consistent with the non-conserved sequence and structure of its loop5 insertion.  相似文献   

16.
The transmission of infectious prions into different host species requires compatible prion protein (PrP) primary structures, and even one heterologous residue at a pivotal position can block prion infection. Mapping the key amino acid positions that govern cross-species prion conversion has not yet been possible, although certain residue positions have been identified as restrictive, including residues in the β22 loop region of PrP. To further define how β22 residues impact conversion, we investigated residue substitutions in PrPC using an in vitro prion conversion assay. Within the β22 loop, a tyrosine residue at position 169 is strictly conserved among mammals, and transgenic mice expressing mouse PrP having the Y169G, S170N, and N174T substitutions resist prion infection. To better understand the structural requirements of specific residues for conversion initiated by mouse prions, we substituted a diverse array of amino acids at position 169 of PrP. We found that the substitution of glycine, leucine, or glutamine at position 169 reduced conversion by ∼75%. In contrast, replacing tyrosine 169 with either of the bulky, aromatic residues, phenylalanine or tryptophan, supported efficient prion conversion. We propose a model based on a requirement for tightly interdigitating complementary amino acid side chains within specific domains of adjacent PrP molecules, known as “steric zippers,” to explain these results. Collectively, these studies suggest that an aromatic residue at position 169 supports efficient prion conversion.  相似文献   

17.
18.
19.
20.
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