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 890 毫秒
1.
Thoden JB  Kappock TJ  Stubbe J  Holden HM 《Biochemistry》1999,38(47):15480-15492
Escherichia coli PurK, a dimeric N5-carboxyaminoimidazole ribonucleotide (N5-CAIR) synthetase, catalyzes the conversion of 5-aminoimidazole ribonucleotide (AIR), ATP, and bicarbonate to N5-CAIR, ADP, and Pi. Crystallization of both a sulfate-liganded and the MgADP-liganded E. coli PurK has resulted in structures at 2.1 and 2.5 A resolution, respectively. PurK belongs to the ATP grasp superfamily of C-N ligase enzymes. Each subunit of PurK is composed of three domains (A, B, and C). The B domain contains a flexible, glycine-rich loop (B loop, T123-G130) that is disordered in the sulfate-PurK structure and becomes ordered in the MgADP-PurK structure. MgADP is wedged between the B and C domains, as with all members of the ATP grasp superfamily. Other enzymes in this superfamily contain a conserved Omega loop proposed to interact with the B loop, define the specificity of their nonnucleotide substrate, and protect the acyl phosphate intermediate formed from this substrate. PurK contains a minimal Omega loop without conserved residues. In the reaction catalyzed by PurK, carboxyphosphate is the putative acyl phosphate intermediate. The sulfate of the sulfate ion-liganded PurK interacts electrostatically with Arg 242 and the backbone amide group of Asn 245, components of the J loop of the C domain. This sulfate may reveal the location of the carboxyphosphate binding site. Conserved residues within the C-terminus of the C domain define a pocket that is proposed to bind AIR in collaboration with an N-terminal strand loop helix motif in the A domain (P loop, G8-L1). The P loop is proposed to bind the phosphate of AIR on the basis of similar binding sites observed in PurN and PurE and proposed in PurD and PurT, four other enzymes in the purine pathway.  相似文献   

2.
In an earlier study, we showed that two‐domain segment‐swapped proteins can evolve by domain swapping and fusion, resulting in a protein with two linkers connecting its domains. We proposed that a potential evolutionary advantage of this topology may be the restriction of interdomain motions, which may facilitate domain closure by a hinge‐like movement, crucial for the function of many enzymes. Here, we test this hypothesis computationally on uroporphyrinogen III synthase, a two‐domain segment‐swapped enzyme essential in porphyrin metabolism. To compare the interdomain flexibility between the wild‐type, segment‐swapped enzyme (having two interdomain linkers) and circular permutants of the same enzyme having only one interdomain linker, we performed geometric and molecular dynamics simulations for these species in their ligand‐free and ligand‐bound forms. We find that in the ligand‐free form, interdomain motions in the wild‐type enzyme are significantly more restricted than they would be with only one interdomain linker, while the flexibility difference is negligible in the ligand‐bound form. We also estimated the entropy costs of ligand binding associated with the interdomain motions, and find that the change in domain connectivity due to segment swapping results in a reduction of this entropy cost, corresponding to ~20% of the total ligand binding free energy. In addition, the restriction of interdomain motions may also help the functional domain‐closure motion required for catalysis. This suggests that the evolution of the segment‐swapped topology facilitated the evolution of enzyme function for this protein by influencing its dynamic properties. Proteins 2016; 85:46–53. © 2016 Wiley Periodicals, Inc.  相似文献   

3.
Aspartate transcarbamoylase and dihydroorotase, enzymes that catalyze the second and third step in de novo pyrimidine biosynthesis, are associated in dodecameric complexes in Aquifex aeolicus and many other organisms. The architecture of the dodecamer is ideally suited to channel the intermediate, carbamoyl aspartate from its site of synthesis on the ATC subunit to the active site of DHO, which catalyzes the next step in the pathway, because both reactions occur within a large, internal solvent‐filled cavity. Channeling usually requires that the reactions of the enzymes are coordinated so that the rate of synthesis of the intermediate matches its rate of utilization. The linkage between the ATC and DHO subunits was demonstrated by showing that the binding of the bisubstrate analog, N‐phosphonacetyl‐L ‐aspartate to the ATC subunit inhibits the activity of the distal DHO subunit. Structural studies identified a DHO loop, loop A, interdigitating between the ATC domains that would be expected to interfere with domain closure essential for ATC catalysis. Mutation of the DHO residues in loop A that penetrate deeply between the two ATC domains inhibits the ATC activity by interfering with the normal reciprocal linkage between the two enzymes. Moreover, a synthetic peptide that mimics that part of the DHO loop that binds between the two ATC domains was found to be an allosteric or noncompletive ATC inhibitor (Ki = 22 μM). A model is proposed suggesting that loop A is an important component of the functional linkage between the enzymes.  相似文献   

4.
Mycobacterium tuberculosis encodes five gene clusters (ESX‐1 to ESX‐5) for Type VII protein secretion systems that are implicated in mycobacterial pathogenicity. Substrates for the secretion apparatus are encoded within the gene clusters and in additional loci that lack the components of the secretion apparatus. The best characterized substrates are the ESX complexes, 1:1 heterodimers of ESAT‐6 and CFP‐10, the prototypical member that has been shown to be essential for Mycobacterium tuberculosis pathogenesis. We have determined the structure of EsxRS, a homolog of EsxGH of the ESX‐3 gene cluster, at 1.91 Å resolution. The EsxRS structure is composed of two four‐helix bundles resulting from the 3D domain swapping of the C‐terminal domain of EsxS, the CFP‐10 homolog. The four‐helix bundles at the extremities of the complex have a similar architecture to the structure of ESAT‐6·CFP‐10 (EsxAB) of ESX‐1, but in EsxRS a hinge loop linking the α‐helical domains of EsxS undergoes a loop‐to‐helix transition that creates the domain swapped EsxRS tetramer. Based on the atomic structure of EsxRS and existing biochemical data on ESX complexes, we propose that higher order ESX oligomers may increase avidity of ESX binding to host receptor molecules or, alternatively, the conformational change that creates the domain swapped structure may be the basis of ESX complex dissociation that would free ESAT‐6 to exert a cytotoxic effect.  相似文献   

5.
The number of artificial protein supramolecules has been increasing; however, control of protein oligomer formation remains challenging. Cytochrome c′ from Allochromatium vinosum (AVCP) is a homodimeric protein in its native form, where its protomer exhibits a four‐helix bundle structure containing a covalently bound five‐coordinate heme as a gas binding site. AVCP exhibits a unique reversible dimer–monomer transition according to the absence and presence of CO. Herein, domain‐swapped dimeric AVCP was constructed and utilized to form a tetramer and high‐order oligomers. The X‐ray crystal structure of oxidized tetrameric AVCP consisted of two monomer subunits and one domain‐swapped dimer subunit, which exchanged the region containing helices αA and αB between protomers. The active site structures of the domain‐swapped dimer subunit and monomer subunits in the tetramer were similar to those of the monomer subunits in the native dimer. The subunit–subunit interactions at the interfaces of the domain‐swapped dimer and monomer subunits in the tetramer were also similar to the subunit–subunit interaction in the native dimer. Reduced tetrameric AVCP dissociated to a domain‐swapped dimer and two monomers upon CO binding. Without monomers, the domain‐swapped dimers formed tetramers, hexamers, and higher‐order oligomers in the absence of CO, whereas the oligomers dissociated to domain‐swapped dimers in the presence of CO, demonstrating that the domain‐swapped dimer maintains the CO‐induced subunit dissociation behavior of native ACVP. These results suggest that protein oligomer formation may be controlled by utilizing domain swapping for a dimer–monomer transition protein.  相似文献   

6.
RNA‐DNA hybrids form throughout the chromosome during normal growth and under stress conditions. When left unresolved, RNA‐DNA hybrids can slow replication fork progression, cause DNA breaks, and increase mutagenesis. To remove hybrids, all organisms use ribonuclease H (RNase H) to specifically degrade the RNA portion. Here we show that, in addition to chromosomally encoded RNase HII and RNase HIII, Bacillus subtilis NCIB 3610 encodes a previously uncharacterized RNase HI protein, RnhP, on the endogenous plasmid pBS32. Like other RNase HI enzymes, RnhP incises Okazaki fragments, ribopatches, and a complementary RNA‐DNA hybrid. We show that while chromosomally encoded RNase HIII is required for pBS32 hyper‐replication, RnhP compensates for the loss of RNase HIII activity on the chromosome. Consequently, loss of RnhP and RNase HIII impairs bacterial growth. We show that the decreased growth rate can be explained by laggard replication fork progression near the terminus region of the right replichore, resulting in SOS induction and inhibition of cell division. We conclude that all three functional RNase H enzymes are present in B. subtilis NCIB 3610 and that the plasmid‐encoded RNase HI contributes to chromosome stability, while the chromosomally encoded RNase HIII is important for chromosome stability and plasmid hyper‐replication.  相似文献   

7.
N‐Acetylneuraminic acid (NANA) is the most common naturally occurring sialic acid and plays a key role in the pathogenesis of a select number of neuroinvasive bacteria such as Neisseria meningitidis. NANA is synthesized in prokaryotes via a condensation reaction between phosphoenolpyruvate and N‐acetylmannosamine. This reaction is catalyzed by a domain swapped, homodimeric enzyme, N‐acetylneuraminic acid synthase (NANAS). NANAS comprises two distinct domains; an N‐terminal catalytic (β/α)8 barrel linked to a C‐terminal antifreeze protein‐like (AFPL) domain. We have investigated the role of the AFPL domain by characterizing a truncated variant of NmeNANAS, which was discovered to be soluble yet inactive. Analytical ultracentrifugation and analytical size exclusion were used to probe the quaternary state of the NmeNANAS truncation, and revealed that loss of the AFPL domain destabilizes the dimeric form of the enzyme. The results from this study thereby demonstrate that the AFPL domain plays a critical role for both the catalytic function and quaternary structure stability of NANAS. Small angle X‐ray scattering, molecular dynamics simulations, and amino acid substitutions expose a complex hydrogen‐bonding relay, which links the roles of the catalytic and AFPL domains across subunit boundaries. Proteins 2014; 82:2054–2066. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
Glycinamide ribonucleotide formyltransferases (GARTs) are part of the de novo purine biosynthetic pathway, catalyzing the direct transfer of a formyl group from the tetrahydrofolate cofactor to the glycinamide ribonucleotide substrate. Despite the low amino acid-sequence identity between the GARTs from Escherichia coli and human, their tertiary structures are superimposable. As part of our functional studies of these enzymes, we have investigated the interchangeability of individual protein fragments or modules between the two enzymes and the functional properties of the resulting hybrids. The modular nature of GART facilitated the creation of combinatorial libraries of chimeras between the Escherichia coli and human enzymes, which were functionally selected through complementation of an auxotrophic Escherichia coli strain. From a pool of several dozen sequence distinct hybrids, six in vivo-functional fusion genes were selected, overexpressed, and purified to homogeneity. The kinetic analysis of these constructs and the comparison of their k(cat) and K(M) values to the parental enzymes suggest that the characteristic kinetic properties from the two parents are "modular encoded" and can be exchanged by domain swapping. The chimeras in general, however, are subject to temperature instability and misfolding; thus, they serve primarily as useful candidates for further rounds of optimization.  相似文献   

9.
In Escherichia coli, the PurT-encoded glycinamide ribonucleotide transformylase, or PurT transformylase, catalyzes an alternative formylation of glycinamide ribonucleotide (GAR) in the de novo pathway for purine biosynthesis. On the basis of amino acid sequence analyses, it is known that the PurT transformylase belongs to the ATP-grasp superfamily of proteins. The common theme among members of this superfamily is a catalytic reaction mechanism that requires ATP and proceeds through an acyl phosphate intermediate. All of the enzymes belonging to the ATP-grasp superfamily are composed of three structural motifs, termed the A-, B-, and C-domains, and in each case, the ATP is wedged between the B- and C-domains. Here we describe two high-resolution X-ray crystallographic structures of PurT transformylase from E. coli: one form complexed with the nonhydrolyzable ATP analogue AMPPNP and the second with bound AMPPNP and GAR. The latter structure is of special significance because it represents the first ternary complex to be determined for a member of the ATP-grasp superfamily involved in purine biosynthesis and as such provides new information about the active site region involved in ribonucleotide binding. Specifically in PurT transformylase, the GAR substrate is anchored to the protein via Glu 82, Asp 286, Lys 355, Arg 362, and Arg 363. Key amino acid side chains involved in binding the AMPPNP to the enzyme include Arg 114, Lys 155, Glu 195, Glu 203, and Glu 267. Strikingly, the amino group of GAR that is formylated during the reaction lies at 2.8 A from one of the gamma-phosphoryl oxygens of the AMPPNP.  相似文献   

10.
11.
The adenosine monoposphate‐forming acyl‐CoA synthetase enzymes catalyze a two‐step reaction that involves the initial formation of an acyl adenylate that reacts in a second partial reaction to form a thioester between the acyl substrate and CoA. These enzymes utilize a Domain Alternation catalytic mechanism, whereby a ~110 residue C‐terminal domain rotates by 140° to form distinct catalytic conformations for the two partial reactions. The structure of an acetoacetyl‐CoA synthetase (AacS) is presented that illustrates a novel aspect of this C‐terminal domain. Specifically, several acetyl‐ and acetoacetyl‐CoA synthetases contain a 30‐residue extension on the C‐terminus compared to other members of this family. Whereas residues from this extension are disordered in prior structures, the AacS structure shows that residues from this extension may interact with key catalytic residues from the N‐terminal domain. Proteins 2015; 83:575–581. © 2014 Wiley Periodicals, Inc.  相似文献   

12.
To recycle reduced sulfur to methionine in the methionine salvage pathway (MSP), 5-methylthioribulose-1-phosphate is converted to 2-keto-4-methylthiobutyrate, the methionine precursor, by four steps; dehydratase, enolase, phosphatase, and dioxygenase reactions (catalyzed by MtnB, MtnW, MtnX and MtnD, respectively, in Bacillus subtilis). It has been proposed that the MtnBD fusion enzyme in Tetrahymena thermophila catalyzes four sequential reactions from the dehydratase to dioxygenase steps, based on the results of molecular biological analyses of mutant yeast strains with knocked-out MSP genes, suggesting that new catalytic function can be acquired by fusion of enzymes. This result raises the question of how the MtnBD fusion enzyme can catalyze four very different reactions, especially since there are no homologous domains for enolase and phosphatase (MtnW and MtnX, respectively, in B. subtilis) in the peptide. Here, we tried to identify the domains responsible for catalyzing the four reactions using recombinant proteins of full-length MtnBD and each domain alone. UV-visible and 1H-NMR spectral analyses of reaction products revealed that the MtnB domain catalyzes dehydration and enolization and the MtnD domain catalyzes dioxygenation. Contrary to a previous report, conversion of 5-methylthioribulose-1-phosphate to 2-keto-4-methylthiobutyrate was dependent on addition of an exogenous phosphatase from B. subtilis. This was observed for both the MtnB domain and full-length MtnBD, suggesting that MtnBD does not catalyze the phosphatase reaction. Our results suggest that the MtnB domain of T. thermophila MtnBD acquired the new function to catalyze both the dehydratase and enolase reactions through evolutionary gene mutations, rather than fusion of MSP genes.  相似文献   

13.
In most organisms, dedicated multiprotein complexes, called exosome or RNA degradosome, carry out RNA degradation and processing. In addition to varying exoribonucleases or endoribonucleases, most of these complexes contain a RNA helicase. In the Gram‐positive bacterium Bacillus subtilis, a RNA degradosome has recently been described; however, no RNA helicase was identified. In this work, we tested the interaction of the four DEAD box RNA helicases encoded in the B. subtilis genome with the RNA degradosome components. One of these helicases, CshA, is able to interact with several of the degradosome proteins, i.e. RNase Y, the polynucleotide phosphorylase, and the glycolytic enzymes enolase and phosphofructokinase. The determination of in vivo protein–protein interactions revealed that CshA is indeed present in a complex with polynucleotide phosphorylase. CshA is composed of two RecA‐like domains that are found in all DEAD box RNA helicases and a C‐terminal domain that is present in some members of this protein family. An analysis of the contribution of the individual domains of CshA revealed that the C‐terminal domain is crucial both for dimerization of CshA and for all interactions with components of the RNA degradosome, including RNase Y. A transfer of this domain to CshB allowed the resulting chimeric protein to interact with RNase Y suggesting that this domain confers interaction specificity. As a degradosome component, CshA is present in the cell in similar amounts under all conditions. Taken together, our results suggest that CshA is the functional equivalent of the RhlB helicase of the Escherichia coli RNA degradosome.  相似文献   

14.
15.
The human genome encodes ~750 G‐protein‐coupled receptors (GPCRs), including prokineticin receptor 2 (PROKR2) involved in the regulation of sexual maturation. Previously reported pathogenic gain‐of‐function mutations of GPCR genes invariably encoded aberrant receptors with excessive signal transduction activity. Although in vitro assays demonstrated that an artificially created inactive mutant of PROKR2 exerted paradoxical gain‐of‐function effects when co‐transfected with wild‐type proteins, such a phenomenon has not been observed in vivo. Here, we report a heterozygous frameshift mutation of PROKR2 identified in a 3.5‐year‐old girl with central precocious puberty. The mutant mRNA escaped nonsense‐mediated decay and generated a GPCR lacking two transmembrane domains and the carboxyl‐terminal tail. The mutant protein had no in vitro signal transduction activity; however, cells co‐expressing the mutant and wild‐type PROKR2 exhibited markedly exaggerated ligand‐induced Ca2+ responses. The results indicate that certain inactive PROKR2 mutants can cause early puberty by enhancing the functional property of coexisting wild‐type proteins. Considering the structural similarity among GPCRs, this paradoxical gain‐of‐function mechanism may underlie various human disorders.  相似文献   

16.
Biosynthesis of asymmetric carotenoids such as α‐carotene and lutein in plants and green algae involves the two enzymes lycopene β‐cyclase (LCYB) and lycopene ε‐cyclase (LCYE). The two cyclases are closely related and probably resulted from an ancient gene duplication. While in most plants investigated so far the two cyclases are encoded by separate genes, prasinophyte algae of the order Mamiellales contain a single gene encoding a fusion protein comprised of LCYB, LCYE and a C‐terminal light‐harvesting complex (LHC) domain. Here we show that the lycopene cyclase fusion protein from Ostreococcus lucimarinus catalyzed the simultaneous formation of α‐carotene and β‐carotene when heterologously expressed in Escherichia coli. The stoichiometry of the two products in E. coli could be altered by gradual truncation of the C‐terminus, suggesting that the LHC domain may be involved in modulating the relative activities of the two cyclase domains in the algae. Partial deletions of the linker region between the cyclase domains or replacement of one or both cyclase domains with the corresponding cyclases from the green alga Chlamydomonas reinhardtii resulted in pronounced shifts of the α‐carotene‐to‐β‐carotene ratio, indicating that both the relative activities of the cyclase domains and the overall structure of the fusion protein have a strong impact on the product stoichiometry. The possibility to tune the product ratio of the lycopene cyclase fusion protein from Mamiellales renders it useful for the biotechnological production of the asymmetric carotenoids α‐carotene or lutein in bacteria or fungi.  相似文献   

17.
The Cu+‐ATPase CopA from Archaeoglobus fulgidus belongs to the P1B family of the P‐type ATPases. These integral membrane proteins couple the energy of ATP hydrolysis to heavy metal ion translocation across membranes. A defining feature of P1B‐1‐type ATPases is the presence of soluble metal binding domains at the N‐terminus (N‐MBDs). The N‐MBDs exhibit a conserved ferredoxin‐like fold, similar to that of soluble copper chaperones, and bind metal ions via a conserved CXXC motif. The N‐MBDs enable Cu+ regulation of turnover rates apparently through Cu‐sensitive interactions with catalytic domains. A. fulgidus CopA is unusual in that it contains both an N‐terminal MBD and a C‐terminal MBD (C‐MBD). The functional role of the unique C‐MBD has not been established. Here, we report the crystal structure of the apo, oxidized C‐MBD to 2.0 Å resolution. In the structure, two C‐MBD monomers form a domain‐swapped dimer, which has not been observed previously for similar domains. In addition, the interaction of the C‐MBD with the other cytoplasmic domains of CopA, the ATP binding domain (ATPBD) and actuator domain (A‐domain), has been investigated. Interestingly, the C‐MBD interacts specifically with both of these domains, independent of the presence of Cu+ or nucleotides. These data reinforce the uniqueness of the C‐MBD and suggest a distinct structural role for the C‐MBD in CopA transport. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

18.
The FK506‐binding protein (FKBP) family consists of proteins with a variety of protein–protein interaction domains and versatile cellular functions. It is assumed that all members are peptidyl‐prolyl cis–trans isomerases with the enzymatic function attributed to the FKBP domain. Six members of this family localize to the mammalian endoplasmic reticulum (ER). Four of them, FKBP22 (encoded by the FKBP14 gene), FKBP23 (FKBP7), FKBP60 (FKBP9), and FKBP65 (FKBP10), are unique among all FKBPs as they contain the EF‐hand motifs. Little is known about the biological roles of these proteins, but emerging genetics studies are attracting great interest to the ER resident FKBPs, as mutations in genes encoding FKBP10 and FKBP14 were shown to cause a variety of matrix disorders. Although the structural organization of the FKBP‐type domain as well as of the EF‐hand motif has been known for a while, it is difficult to conclude how these structures are combined and how it affects the protein functionality. We have determined a unique 1.9 Å resolution crystal structure for human FKBP22, which can serve as a prototype for other EF hand‐containing FKBPs. The EF‐hand motifs of two FKBP22 molecules form a dimeric complex with an elongated and predominantly hydrophobic cavity that can potentially be occupied by an aliphatic ligand. The FKBP‐type domains are separated by a cleft and their putative active sites can catalyze isomerazation of two bonds within a polypeptide chain in extended conformation. These structural results are of prime interest for understanding biological functions of ER resident FKBPs containing EF‐hand motifs.  相似文献   

19.

Background  

Non-ribosomal peptide synthetases (NRPSs) are large multimodular enzymes that synthesize a wide range of biologically active natural peptide compounds, of which many are pharmacologically important. Peptide bond formation is catalyzed by the Condensation (C) domain. Various functional subtypes of the C domain exist: AnLCL domain catalyzes a peptide bond between two L-amino acids, aDCL domain links an L-amino acid to a growing peptide ending with a D-amino acid, a Starter C domain (first denominated and classified as a separate subtype here) acylates the first amino acid with a β -hydroxy-carboxylic acid (typically a β -hydroxyl fatty acid), and Heterocyclization (Cyc) domains catalyze both peptide bond formation and subsequent cyclization of cysteine, serine or threonine residues. The homologous Epimerization (E) domain flips the chirality of the last amino acid in the growing peptide; Dual E/C domains catalyze both epimerization and condensation.  相似文献   

20.
The multi‐subunit Ca2+/calmodulin‐dependent protein kinase II (CaMKII) holoenzyme plays a critical role in animal learning and memory. The kinase domain of CaMKII is connected by a flexible linker to a C‐terminal hub domain that assembles into a 12‐ or 14‐subunit scaffold that displays the kinase domains around it. Studies on CaMKII suggest that the stoichiometry and dynamic assembly/disassembly of hub oligomers may be important for CaMKII regulation. Although CaMKII is a metazoan protein, genes encoding predicted CaMKII‐like hub domains, without associated kinase domains, are found in the genomes of some green plants and bacteria. We show that the hub domains encoded by three related green algae, Chlamydomonas reinhardtii, Volvox carteri f. nagarensis, and Gonium pectoral, assemble into 16‐, 18‐, and 20‐subunit oligomers, as assayed by native protein mass spectrometry. These are the largest known CaMKII hub domain assemblies. A crystal structure of the hub domain from C. reinhardtii reveals an 18‐subunit organization. We identified four intra‐subunit hydrogen bonds in the core of the fold that are present in the Chlamydomonas hub domain, but not in metazoan hubs. When six point mutations designed to recapitulate these hydrogen bonds were introduced into the human CaMKII‐α hub domain, the mutant protein formed assemblies with 14 and 16 subunits, instead of the normal 12‐ and 14‐subunit assemblies. Our results show that the stoichiometric balance of CaMKII hub assemblies can be shifted readily by small changes in sequence.  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