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 131 毫秒
1.
In Escherichia coli, penicillin-binding protein 3 (PBP3), also known as FtsI, is a central component of the divisome, catalyzing cross-linking of the cell wall peptidoglycan during cell division. PBP3 is mainly periplasmic, with a 23 residues cytoplasmic tail and a single transmembrane helix. We have solved the crystal structure of a soluble form of PBP3 (PBP357–577) at 2.5 Å revealing the two modules of high molecular weight class B PBPs, a carboxy terminal module exhibiting transpeptidase activity and an amino terminal module of unknown function. To gain additional insight, the PBP3 Val88-Ser165 subdomain (PBP388–165), for which the electron density is poorly defined in the PBP3 crystal, was produced and its structure solved by SAD phasing at 2.1 Å. The structure shows a three dimensional domain swapping with a β-strand of one molecule inserted between two strands of the paired molecule, suggesting a possible role in PBP357–577 dimerization.  相似文献   

2.
Bacteriorhodopsins are a large family of seven-helical transmembrane proteins that function as light-driven proton pumps. Here, we present the crystal structure of a new member of the family, Haloarcula marismortui bacteriorhodopsin I (HmBRI) D94N mutant, at the resolution of 2.5 Å. While the HmBRI retinal-binding pocket and proton donor site are similar to those of other archaeal proton pumps, its proton release region is extended and contains additional water molecules. The protein''s fold is reinforced by three novel inter-helical hydrogen bonds, two of which result from double substitutions relative to Halobacterium salinarum bacteriorhodopsin and other similar proteins. Despite the expression in Escherichia coli and consequent absence of native lipids, the protein assembles as a trimer in crystals. The unique extended loop between the helices D and E of HmBRI makes contacts with the adjacent protomer and appears to stabilize the interface. Many lipidic hydrophobic tail groups are discernible in the membrane region, and their positions are similar to those of archaeal isoprenoid lipids in the crystals of other proton pumps, isolated from native or native-like sources. All these features might explain the HmBRI properties and establish the protein as a novel model for the microbial rhodopsin proton pumping studies.  相似文献   

3.
Anaphase-promoting complex or cyclosome (APC/C) is an unusual E3 ubiquitin ligase and an essential protein that controls mitotic progression. APC/C includes at least 13 subunits, but no structure has been determined for any tetratricopeptide repeat (TPR)-containing subunit (Apc3 and -6-8) in the TPR subcomplex of APC/C. Apc7 is a TPR-containing subunit that exists only in vertebrate APC/C. Here we report the crystal structure of quad mutant of nApc7 (N-terminal fragment, residues 1-147) of human Apc7 at a resolution of 2.5 Å. The structure of nApc7 adopts a TPR-like motif and has a unique dimerization interface, although the protein does not contain the conserved TPR sequence. Based on the structure of nApc7, in addition to previous experimental findings, we proposed a putative homodimeric structure for full-length Apc7. This model suggests that TPR-containing subunits self-associate and bind to adaptors and substrates via an IR peptide in TPR-containing subunits of APC/C.Anaphase-promoting complex/cyclosome (APC/C)2 is an E3 ubiquitin ligase that controls mitotic progression (1). APC/C is an ∼1.7-MDa protein complex that is composed of at least 13 subunits, and it contains a cullin homolog (Apc2), a ring-H2 finger domain (Apc11), and a tetratricopeptide repeat (TPR)-containing subunit (TPR subunit; Apc3 and -6-8) (2). Most TPR subunits are essential and evolutionarily conserved in eukaryotes (3).APC/C requires two adaptors that contain a C-terminal WD40 domain, Cdc20 and Cdh1, to recruit and select various substrates at different stages of the cell cycle. Moreover, both adaptors and specific APC/C subunits contribute to substrate recognition (4).APC/C specifically ubiquitinates cell cycle regulatory proteins that contain destruction (D) or KEN box motifs (5-7), which target them for destruction by the 26 S proteosome (8). During the cell cycle, APC/C mediates the metaphase-anaphase transition by ubiquitinating and degrading securin, a separase inhibitor, which participates in the degradation of chromatic cohesion complexes and ubiquitinates B-type cyclin, thereby accelerating transition from the late mitotic phase to G1 (9). In addition to its primary role in cell cycle regulation, APC/C participates in postmitotic processes, such as regulation of synaptic size and axon growth (10, 11).To assess the mechanism that underlies cell cycle regulation by APC/C and the various roles of its subunits, we need to understand how APC/C is organized into higher order structures and the manner in which the subunits assemble. Although little is known regarding the crystal structures of APC/C components, three-dimensional models of APC/C have recently been obtained by cryo-negative staining EM in human, Xenopus laevis, Saccharomyces cerevisiae, and Schizosaccharomyces pombe (12-15). Several studies have indicated that APC/C assumes an asymmetric triangular shape that is composed of an outer shell and a cavity that extends through its center (12, 14). Furthermore, APC/C includes a catalytic subcomplex (Doc1/Apc10, Apc11, and Apc2), a structural complex (Apc1, Apc4, and Apc5), and a TPR subcomplex (TPR-containing subunits and nonessential subunits) (16).A TPR unit consists of a 34-residue repeat motif that adopts a helix-turn-helix conformation, which is associated with protein-protein interactions (17). Multiple copies of TPR-containing subunits are organized into the TPR subcomplex within APC/C, and this subcomplex is functionally important for the recruitment of adaptors and substrates (18). In fact, adaptors (Cdc20 and Cdh1) and Doc1/Apc10 bind to the C-terminal domain of the TPR-containing subunits Apc3 and Apc7 via the IR peptide tail sequence (7, 16, 19). It is unknown, however, how TPR-containing subunits form homo- and heterosubunit complexes, although studies have demonstrated that TPR-containing subunits self-associate in vivo and in vitro (15) and that they interact with other TPR-containing subunits (20).Apc7 is found only in vertebrate APC/C and is estimated to contain 9-15 TPR motifs, similar to other TPR-containing subunits (9). Apc7 is considered to be a molecular descendant of the same ancestral protein that gave rise to Apc3. Furthermore, the N-terminal domain of Apc7 has been reported to contain cell cycle-regulated phosphorylation sites (21), and the C-terminal TPR domain of Apc7 interacts with Cdh1 and Cdc20 (19). In Drosophila APC/C, the homolog of vertebrate Apc7 participates in synergistic genetic interactions with other TPR-containing subunits (22).The function of Apc7 within vertebrate APC/C, however, is poorly understood. Moreover, although the C-terminal regions of Apc3 and Apc7 include a tandem of nine TPR motifs, the N-terminal domains of human Apc3 and Apc7 share little homology with the canonical TPR sequence. Therefore, the N-terminal domain of human Apc7 is expected to have a significant function in vertebrate APC/C.In this study, we determined the crystal structure of the N-terminal fragment of human Apc7 (residues 1-147, denoted nApc7), and the homodimeric self-association of nApc7 structure led us to insights into mechanisms of vertebrate APC/C.  相似文献   

4.
5.
The subunit molecular mass of glyceraldehyde-3-phosphate dehydrogenase from the extreme halophile Haloarcula vallismortis (hGAPDH) was determined by mass spectrometry to be 35990 +/- 80 daltons, similar to other GAPDHs. Complementary density, sedimentation and light scattering experiments showed the protein to be a tetramer that binds 0.18 +/- 0.10 gram of water and 0.07 +/- 0.02 gram of KCl per gram of protein, in multimolar KCl solutions. At low salt (below 1 M), the tetramer dissociated into unfolded monomers. This is the third halophilic protein for which solvent interactions were measured. The extent of these interactions depends on the protein, but all form an invariant particle, in multimolar NaCl or KCl solutions, that binds a high proportion of salt when compared to non-halophilic proteins.  相似文献   

6.
IQGAP1 is a 190-kDa molecular scaffold containing several domains required for interaction with numerous proteins. One domain is homologous to Ras GTPase-activating protein (GAP) domains. However, instead of accelerating hydrolysis of bound GTP on Ras IQGAP1, using its GAP-related domain (GRD) binds to Cdc42 and Rac1 and stabilizes their GTP-bound states. We report here the crystal structure of the isolated IQGAP1 GRD. Despite low sequence conservation, the overall structure of the GRD is very similar to the GAP domains from p120 RasGAP, neurofibromin, and SynGAP. However, instead of the catalytic “arginine finger” seen in functional Ras GAPs, the GRD has a conserved threonine residue. GRD residues 1099–1129 have no structural equivalent in RasGAP and are seen to form an extension at one end of the molecule. Because the sequence of these residues is highly conserved, this region likely confers a functionality particular to IQGAP family GRDs. We have used isothermal titration calorimetry to demonstrate that the isolated GRD binds to active Cdc42. Assuming a mode of interaction similar to that displayed in the Ras-RasGAP complex, we created an energy-minimized model of Cdc42·GTP bound to the GRD. Residues of the GRD that contact Cdc42 map to the surface of the GRD that displays the highest level of sequence conservation. The model indicates that steric clash between threonine 1046 with the phosphate-binding loop and other subtle changes would likely disrupt the proper geometry required for GTP hydrolysis.The small GTPase Ras functions as a binary switch in cell signaling processes. When bound to GTP, Ras is able to interact with effector proteins, including Raf kinase, and alter their activities. Ras signaling is terminated when bound GTP is hydrolyzed to GDP and inorganic phosphate. The basal rate of GTP hydrolysis on Ras is quite slow (∼1.2 × 10–4 s–1), but this rate of hydrolysis can be enhanced ∼105-fold by interaction with a GTPase-activating protein (GAP)2 (1). Several RasGAPs have been identified to date including p120 RasGAP and neurofibromin (NF1). The Rho family of Ras-related small GTPases also function as binary switches in cell signaling processes. Whereas the intrinsic rate of GTP hydrolysis on Rho proteins is faster than Ras, this rate can also be stimulated by interaction with a RhoGAP. Examination of the structures of the GAP domains of p120RasGAP (2), neurofibromin (3), SynGAP (4), and the GAP domains from the RhoGAPs p50 RhoGAP and the Bcr homology domain of phosphatidylinositol 3-kinase (5, 6) indicates that although ostensibly different, these all-helical domains are structurally related (7).IQGAP1 was discovered by chance during an attempt to isolate novel matrix metalloproteinases (8). Analysis reveals that the protein contains several discrete domains and motifs including a region containing four isoleucine- and glutamine-rich motifs (IQ repeats) and a region with sequence homology to the Ras-specific GAP domains of p120RasGAP, NF1, and SynGAP (24, 8). Subsequently, two homologs, IQGAP2 and IQGAP3, have been discovered. The IQ repeats have been shown to mediate binding to calmodulin and calmodulin-like proteins (e.g. S100, myosin essential light chain), whereas the GAP-related domain (GRD) does not appear to bind to Ras but instead is necessary for binding to the Rho family GTPases Cdc42 and Rac1, primarily in their active forms (911). However, instead of accelerating hydrolysis of GTP, IQGAP1 preserves the activated states of Cdc42 and Rac1 to the extent that overexpression of IQGAP1 in cells increases the levels of active GTPase (12). Because IQGAP1 expression increases the level of activated Cdc42, initially there was some confusion as to whether the protein might not represent a novel guanine nucleotide exchange factor. However it now appears that IQGAP1 is an effector of Cdc42 and Rac1 and preserves their activated states by tightly binding to the GTPases and stabilizing them in a conformation not conducive to GTP hydrolysis. IQGAP1 appears to be such an important effector for Cdc42 that abrogation of binding to IQGAP1 not only reduces the levels of active Cdc42, it also reduces membrane-localized Cdc42 and the cellular response to bradykinin (12).A growing body of evidence implicates IQGAP1 in carcinogenesis. Expression of IQGAP1 increases during the transition from a minimally to a highly metastastic form of melanoma, and IQGAP1 has been found to be overexpressed in ovarian, breast, lung, and colorectal cancers (1317). In vitro, overexpressed IQGAP1 enhances cell motility and invasiveness in a process that requires Cdc42 and Rac (18). β-Catenin is one of the many binding partners of IQGAP1 identified to date. IQGAP1 has been shown to bind to β-catenin and interfere with β-catenin binding to α-catenin, an interaction necessary for stable cell-cell adhesion (19). Another study found that IQGAP2 knock-out mice overexpress IQGAP1 and developage-dependent liver cancer and apoptosis (20).To better understand how a protein domain homologous to others that accelerate GTP hydrolysis can function as an effector and preserve the GTP-bound state, we have determined the x-ray structure of the IQGAP1 GRD. Despite low sequence identity, the GRD structure is quite similar to the GAP domains of p120, neurofibromin, and SynGAP; however, unlike those domains, the GRD possesses a conserved threonine in place of the catalytic arginine finger and has a 31-residue insertion that projects from one end of the molecule. Using the coordinates of Ras·GDP·AlF3 in complex with the GAP domain of p120, we built a model of Cdc42·GTP bound to the GRD. The model indicates that a steric clash between the conserved Thr1046 and the phosphate-binding loop of Cdc42 and other subtle changes within the active site would likely preclude nucleotide hydrolysis. Sequence conservation mapped to the surface of the GRD indicates that the surface with the highest degree of conservation overlaps with the surface that makes contacts to Cdc42 in the model.  相似文献   

7.
TyrA proteins belong to a family of dehydrogenases that are dedicated to l-tyrosine biosynthesis. The three TyrA subclasses are distinguished by their substrate specificities, namely the prephenate dehydrogenases, the arogenate dehydrogenases, and the cyclohexadienyl dehydrogenases, which utilize prephenate, l-arogenate, or both substrates, respectively. The molecular mechanism responsible for TyrA substrate selectivity and regulation is unknown. To further our understanding of TyrA-catalyzed reactions, we have determined the crystal structures of Aquifex aeolicus prephenate dehydrogenase bound with NAD+ plus either 4-hydroxyphenylpyuvate, 4-hydroxyphenylpropionate, or l-tyrosine and have used these structures as guides to target active site residues for site-directed mutagenesis. From a combination of mutational and structural analyses, we have demonstrated that His-147 and Arg-250 are key catalytic and binding groups, respectively, and Ser-126 participates in both catalysis and substrate binding through the ligand 4-hydroxyl group. The crystal structure revealed that tyrosine, a known inhibitor, binds directly to the active site of the enzyme and not to an allosteric site. The most interesting finding though, is that mutating His-217 relieved the inhibitory effect of tyrosine on A. aeolicus prephenate dehydrogenase. The identification of a tyrosine-insensitive mutant provides a novel avenue for designing an unregulated enzyme for application in metabolic engineering.Tyrosine serves as a precursor for the synthesis of proteins and secondary metabolites such as quinones (1-3), alkaloids (4), flavonoids (5), and phenolic compounds (5, 6). In prokaryotes and plants, these compounds are important for viability and normal development (7).The TyrA protein family consists of dehydrogenase homologues that are dedicated to the biosynthesis of l-tyrosine. These enzymes participate in two independent metabolic branches that result in the conversion of prephenate to l-tyrosine, namely the arogenate route and the 4-hydroxyphenylpyruvate (HPP)3 routes. Although both of these pathways utilize a common precursor and converge to produce a common end-product, they differ in the sequential order of enzymatic steps. Through the HPP route, prephenate is first decarboxylated by prephenate dehydrogenase (PD) to yield HPP, which is subsequently transaminated to l-tyrosine via a TyrB homologue (8). Alternatively, through the arogenate route, prephenate is first transaminated to l-arogenate by prephenate aminotransferase and then decarboxylated by arogenate dehydrogenase (AD) to yield l-tyrosine (9-11) (see Fig. 1A).Open in a separate windowFIGURE 1.A, metabolic routes from chorismate leading to the synthesis of l-tyrosine and l-phenylalanine. In the arogenate, 4-hydroxyphenylpyruvate, or phenylpyruvate route, prephenate and arogenate are branch point intermediates in both l-tyrosine and l-phenylalanine biosynthesis. Prephenate dehydrogenase catalyzes the oxidative decarboxylation of prephenate with NAD+ to produce hydroxyphenylpyruvate, NADH, and CO2 (40). B, a comparison of the chemical structure of the three ligands, HPP, HPpropionate, and tyrosine, used in the crystallization of A. aeolicus prephenate dehydrogenase. These ligands all have an -OH at the C4 position and a propionyl side chain at the C1 position of the ring.There are three classes of TyrA enzymes that catalyze the oxidative decarboxylation reactions in these two pathways. The enzymes are distinguished by the affinity for cyclohexadienyl substrates. PD and AD accept prephenate or l-arogenate, respectively, whereas the cyclohexadienyl dehydrogenases can catalyze the reaction using either substrate (12).To ensure efficient metabolite distribution of the pathway intermediates, TyrA enzymes are highly regulated by various control mechanisms, including feedback inhibition, and genetic regulation by the Tyr operon (13-16). In some cases, l-tyrosine competes directly with substrate, be it prephenate or l-arogenate for the active site of arogenate or cyclohexadienyl dehydrogenases (14, 17-19). The product HPP can also serve as an efficient competitive inhibitor with respect to prephenate (20). Additionally, at the protein level PDs are only shown to be regulated at distinct allosteric sites or domains to modulate their activity. For example, the results of kinetic studies on the bifunctional Escherichia coli chorismate mutase-prephenate dehydrogenase (CM-PD) have indicated that this enzyme likely possesses a distinct allosteric site for binding tyrosine (21). In contrast, the Bacillus subtilis PD is the only enzyme reported to be competitively inhibited by HPP and l-tyrosine but is also noncompetitively inhibited by l-phenylalanine and l-tryptophan (12, 22). Additional regulatory control is thought to originate through a C-terminal aspartate kinase-CM-TyrA domain of the B. subtilis PD (23).Biochemical analyses of PD from E. coli CM-PD have provided a framework for understanding the molecular mechanism of the TyrA enzymes. The E. coli PD-catalyzed reaction proceeds though a rapid equilibrium, random kinetic mechanism with catalysis as the rate-limiting step (24). Additionally, studies of the pH dependence of the kinetic parameters V and V/K indicate that a deprotonated group facilitates hydride transfer from prephenate to NAD+ by polarizing the 4-hydroxyl group of prephenate, whereas a protonated residue is required for binding prephenate to the enzyme·NAD+ complex (25). The conserved residues His-197 and Arg-294 have been identified through extensive mutagenesis studies to fulfill these two roles (26, 27). Further analyses of the activities of wild-type protein and site-directed variants in the presence of a series of inhibitory substrate analogues support the idea that Arg-294 binds prephenate through the ring carboxylate (26).The structures of AD from Synechocystis sp. and PD from Aquifex aeolicus (both in complex with NAD+) have been reported by Legrand et al. (28) and by our group (29), respectively. Analyses of these structures have provided structural information on the conserved histidine and arginine residues. The structure A. aeolicus PD has also led to the identification of other active site residues that may play a role in enzyme catalysis, most notably Ser-126, which we propose facilitates catalysis by orienting the catalytic histidine and the nicotinamide moiety of NAD+ into their catalytically efficient conformations. Ambiguities can arise from examination of the binary complexes, because prephenate has only been modeled in the active site. For example, analysis of the AD structure by Legrand et al. (28) places Arg-217 (equivalent to Arg-294 in E. coli and Arg-250 in A. aeolicus) too far from the active site to play a role in prephenate binding. Thus, the full complement of interactions between prephenate and TyrA proteins are still largely unknown, as are the interactions of the enzymes with l-tyrosine.To further investigate the importance of residues involved in ligand binding, specificity, and catalysis, we have carried out co-crystallization studies of A. aeolicus PD with NAD+ and prephenate, with NAD+ and 4-hydroxyphenylpropionate (HPpropionate), a product analogue, and with NAD+ and l-tyrosine. Accordingly, this study provides the first direct evidence that l-tyrosine binds to the active site of a prephenate dehydrogenase. We have investigated the role of Ser-126, His-147, His-217, and Arg-250 through the kinetic analysis of site-directed mutants and structural analysis of the co-crystal complexes. To understand the role of active site residues in substrate selectivity, comparative structural analysis of AD and PD was also conducted. The current study provides a basis for understanding the mechanism of substrate selectivity between the different classes of TyrA enzymes and details how A. aeolicus PD can accept prephenate as substrate and l-tyrosine as a competitive inhibitor.  相似文献   

8.
9.
Photosystem II (PSII) catalyzes light-induced water splitting, leading to the evolution of molecular oxygen indispensible for life on the earth. The crystal structure of PSII from cyanobacteria has been solved at an atomic level, but the structure of eukaryotic PSII has not been analyzed. Because eukaryotic PSII possesses additional subunits not found in cyanobacterial PSII, it is important to solve the structure of eukaryotic PSII to elucidate their detailed functions, as well as evolutionary relationships. Here we report the structure of PSII from a red alga Cyanidium caldarium at 2.76 Å resolution, which revealed the structure and interaction sites of PsbQ′, a unique, fourth extrinsic protein required for stabilizing the oxygen-evolving complex in the lumenal surface of PSII. The PsbQ′ subunit was found to be located underneath CP43 in the vicinity of PsbV, and its structure is characterized by a bundle of four up-down helices arranged in a similar way to those of cyanobacterial and higher plant PsbQ, although helices I and II of PsbQ′ were kinked relative to its higher plant counterpart because of its interactions with CP43. Furthermore, two novel transmembrane helices were found in the red algal PSII that are not present in cyanobacterial PSII; one of these helices may correspond to PsbW found only in eukaryotic PSII. The present results represent the first crystal structure of PSII from eukaryotic oxygenic organisms, which were discussed in comparison with the structure of cyanobacterial PSII.  相似文献   

10.
11.
Raman microscopy permits structural analysis of protein crystals in situ in hanging drops, allowing for comparison with Raman measurements in solution. Nevertheless, the two methods sometimes reveal subtle differences in structure that are often ascribed to the water layer surrounding the protein. The novel method of drop-coating deposition Raman spectropscopy (DCDR) exploits an intermediate phase that, although nominally “dry,” has been shown to preserve protein structural features present in solution. The potential of this new approach to bridge the structural gap between proteins in solution and in crystals is explored here with extrinsic protein PsbP of photosystem II from Spinacia oleracea. In the high-resolution (1.98 Å) x-ray crystal structure of PsbP reported here, several segments of the protein chain are present but unresolved. Analysis of the three kinds of Raman spectra of PsbP suggests that most of the subtle differences can indeed be attributed to the water envelope, which is shown here to have a similar Raman intensity in glassy and crystal states. Using molecular dynamics simulations cross-validated by Raman solution data, two unresolved segments of the PsbP crystal structure were modeled as loops, and the amino terminus was inferred to contain an additional beta segment. The complete PsbP structure was compared with that of the PsbP-like protein CyanoP, which plays a more peripheral role in photosystem II function. The comparison suggests possible interaction surfaces of PsbP with higher-plant photosystem II. This work provides the first complete structural picture of this key protein, and it represents the first systematic comparison of Raman data from solution, glassy, and crystalline states of a protein.  相似文献   

12.
The prokaryotic immune system CRISPR/Cas (Clustered Regularly Interspaced Short Palindromic Repeats/CRISPR-associated genes) adapts to foreign invaders by acquiring their short deoxyribonucleic acid (DNA) fragments as spacers, which guide subsequent interference to foreign nucleic acids based on sequence matching. The adaptation mechanism avoiding acquiring ‘self’ DNA fragments is poorly understood. In Haloarcula hispanica, we previously showed that CRISPR adaptation requires being primed by a pre-existing spacer partially matching the invader DNA. Here, we further demonstrate that flanking a fully-matched target sequence, a functional PAM (protospacer adjacent motif) is still required to prime adaptation. Interestingly, interference utilizes only four PAM sequences, whereas adaptation-priming tolerates as many as 23 PAM sequences. This relaxed PAM selectivity explains how adaptation-priming maximizes its tolerance of PAM mutations (that escape interference) while avoiding mis-targeting the spacer DNA within CRISPR locus. We propose that the primed adaptation, which hitches and cooperates with the interference pathway, distinguishes target from non-target by CRISPR ribonucleic acid guidance and PAM recognition.  相似文献   

13.
Tandem repeat (TR) regions are common in yeast adhesins, but their structures are unknown, and their activities are poorly understood. TR regions in Candida albicans Als proteins are conserved glycosylated 36-residue sequences with cell-cell aggregation activity (J. M. Rauceo, R. De Armond, H. Otoo, P. C. Kahn, S. A. Klotz, N. K. Gaur, and P. N. Lipke, Eukaryot. Cell 5:1664–1673, 2006). Ab initio modeling with either Rosetta or LINUS generated consistent structures of three-stranded antiparallel β-sheet domains, whereas randomly shuffled sequences with the same composition generated various structures with consistently higher energies. O- and N-glycosylation patterns showed that each TR domain had exposed hydrophobic surfaces surrounded by glycosylation sites. These structures are consistent with domain dimensions and stability measurements by atomic force microscopy (D. Alsteen, V. Dupres, S. A. Klotz, N. K. Gaur, P. N. Lipke, and Y. F. Dufrene, ACS Nano 3:1677–1682, 2009) and with circular dichroism determination of secondary structure and thermal stability. Functional assays showed that the hydrophobic surfaces of TR domains supported binding to polystyrene surfaces and other TR domains, leading to nonsaturable homophilic binding. The domain structures are like “classic” subunit interaction surfaces and can explain previously observed patterns of promiscuous interactions between TR domains in any Als proteins or between TR domains and surfaces of other proteins. Together, the modeling techniques and the supporting data lead to an approach that relates structure and function in many kinds of repeat domains in fungal adhesins.Yeast adhesins are a diverse set of cell adhesion proteins that mediate adhesion to host cells, environmental substrates, other fungi, and coinfecting bacteria (6, 8, 20, 21, 23, 29). The adhesins share common features, including compact N-terminal domains similar to Ig or lectin domains, Thr-rich midpieces, often in tandem repeats, and long highly glycosylated Ser/Thr-rich C-terminal regions that extend the functional domains out from the cell surface. No structures for the Thr-rich midpieces are known, but they can mediate aggregation of fungal cells (33, 35, 47). The prevalence and conservation of such repeats argue that they are functionally important, despite limited data on their structure and function.In Candida albicans, the Als adhesins are homologous proteins, products of 8 loci that encode numerous alleles of cell surface adhesins (16). In each mature Als protein, there are, from the N terminus, three tandem Ig-like domains, a β-sheet-rich conserved 127-residue amyloid-forming T region, a variable number of 36-residue tandem repeats (TRs), and a highly glycosylated stalk region that extends the N-terminal domains away from the cell surface (Fig. 1) (16, 33, 41). The C termini of these and other wall-associated adhesins are covalently cross-linked into the cell wall through transglycosylation of a modified glycosylphosphatidylinositol (GPI) anchor (18, 25). This modular design, including tandem repeats, is typical of fungal adhesins (8).Open in a separate windowFig. 1.Schematic diagram of the sequence of Als5p. The regions are named above, and the number of amino acid residues in each region is shown below. The modeled sequences are in the TR region.The Als protein Ig-like region, T region, and TR region all have protein-protein interaction activities (26, 33, 35). The Ig-like regions can interact with diverse mammalian proteins, presumably in a way analogous to antibody-antigen binding, as has been shown in the homologous protein α-agglutinin from Saccharomyces cerevisiae (8, 24, 26, 35). The T regions interact through formation of amyloid-like structures both in vivo and in vitro (33, 34a, 36). An insight into the function of the tandem repeats followed from observations that Als proteins initiate and maintain cell-to-cell aggregations, either spontaneously (“autoaggregation”) or following adhesion to a bead-bound defined ligand (10, 11, 36). Aggregation is more extensive for Als proteins with more tandem repeats (26, 35). This result suggested that the tandem repeats are uniquely structured to facilitate or mediate the aggregative function. Circular dichroism spectroscopy of the TR region of Als5p shows a β-sheet-rich structure in the soluble protein (35).In support of their direct involvement in aggregation, the repeat region of the C. albicans adhesin Als5p mediates cell-cell aggregation in the absence of the Ig-like and T domains (35). Moreover, the repeats can also potentiate binding of Als5p to fibronectin (35). Thus, the TR domains mediate cellular aggregation and increased binding to fibronectin. In addition, TR domains and their amino acid sequences are highly conserved across several Candida species (3). These properties need to be explained by their three-dimensional structure.Because there are no homologous structures known, we modeled by two independent ab initio methods. Rosetta assembles structures by combining short peptide structures extracted from the protein structural database PDB (38), then combines structures in a Monte Carlo approach, and assesses energetics of assembled structures. Rosetta has recently been shown to generate accurate models for protein-sized domains (40). We also predicted structures with LINUS, which generates randomized structures and rapidly estimates energetics to choose low-energy models (45). The models were supported by structural analyses with atomic force microscopy and circular dichroism spectroscopy. Functional assays showed that the TR domains can mediate binding activities predicted from the calculated structures.  相似文献   

14.
Urease, a nickel-dependent metalloenzyme, is synthesized by plants, some bacteria, and fungi. It catalyzes the hydrolysis of urea into ammonia and carbon dioxide. Although the amino acid sequences of plant and bacterial ureases are closely related, some biological activities differ significantly. Plant ureases but not bacterial ureases possess insecticidal properties independent of its ureolytic activity. To date, the structural information is available only for bacterial ureases although the jack bean urease (Canavalia ensiformis; JBU), the best-studied plant urease, was the first enzyme to be crystallized in 1926. To better understand the biological properties of plant ureases including the mechanism of insecticidal activity, we initiated the structural studies on some of them. Here, we report the crystal structure of JBU, the first plant urease structure, at 2.05 Å resolution. The active-site architecture of JBU is similar to that of bacterial ureases containing a bi-nickel center. JBU has a bound phosphate and covalently modified residue (Cys592) by β-mercaptoethanol at its active site, and the concomitant binding of multiple inhibitors (phosphate and β-mercaptoethanol) is not observed so far in bacterial ureases. By correlating the structural information of JBU with the available biophysical and biochemical data on insecticidal properties of plant ureases, we hypothesize that the amphipathic β-hairpin located in the entomotoxic peptide region of plant ureases might form a membrane insertion β-barrel as found in β-pore-forming toxins.  相似文献   

15.
3-Deazauridine: Crystal Structure and Conformation   总被引:1,自引:0,他引:1  
THE nucleoside analogue 3-deazauridine1,2 (Fig. 1), 4-hydroxy-1-(β-D-ribo-pentafuranosyl)-2-pyridone, has powerful cytostatic properties; at a concentration of 5 × 10?6 M it inhibits the growth of Ehrlich ascites cells by 50% (ref. 3). In the codon-specific binding of tRNAphe to ribosomes, 3-deazauridine can substitute for uridine in the 5′ - and middle position of the codon but not in the wobble position4. NMR evidence1,2 suggests that the major form in solution has a 4-hydroxyl group but a dynamic enol-keto tautomerism is taking place. The present X-ray diffraction study was undertaken to determine the tautomeric form present in the crystalline state and to investigate the hydrogen bonding possibilities of 3-deazauridine. We find the enol form with an unusually short hydrogen bond between the 4-hydroxyl group and the 2-keto oxygen of a neighbouring molecule.  相似文献   

16.
Deltarhodopsin, a new member of the microbial rhodopsin family, functions as a light‐driven proton pump. Here, we report the three‐dimensional structure of deltarhodopsin (dR3) from Haloterrigena thermotolerans at 2.7 Å resolution. A crystal belonging to space group R32 (a, b = 111.71 Å, c = 198.25 Å) was obtained by the membrane fusion method. In this crystal, dR3 forms a trimeric structure as observed for bacteriorhodopsin (bR). Structural comparison of dR with bR showed that the inner part (the proton release and uptake pathways) is highly conserved. Meanwhile, residues in the protein–protein contact region are largely altered so that the diameter of the trimeric structure at the cytoplasmic side is noticeably larger in dR3. Unlike bR, dR3 possesses a helical segment at the C‐terminal region that fills the space between the AB and EF loops. A significant difference is also seen in the FG loop, which is one residue longer in dR3. Another peculiar property of dR3 is a highly crowded distribution of positively charged residues on the cytoplasmic surface, which may be relevant to a specific interaction with some cytoplasmic component.Proteins 2013; © 2013 Wiley Periodicals, Inc.  相似文献   

17.
Glyceraldehyde 3-phosphate dehydrogenase (EC 1.2.1.12) from the extremely halophilic archaebacterium Haloarcula vallismortis has been purified in a four step procedure to electrophoretic homogeneity. The enzyme is a tetramer with a relative molecular mass of 160000. It is strictly NAD+-dependent and exhibits its highest activity in 2 mol/l KCl at 45°C. Amino acid analysis and isoelectric focusing indicate an excess of acidic amino acids. Two parts of the primary sequence are reported. These peptides have been compared with glyceraldehyde 3-phosphate dehydrogenases from other archaebacteria, eubacteria and eucaryotes. The peptides show a high grade of similarity to glyceraldehyde 3-phosphate dehydrogenase from eucaryotes.Abbreviations BCA bicinchoninic acid - CTAB cetyltrimethyl ammonium bromide - DTE dithioerythritol - DTT dithiothreitol - GAP glyccraldehyde 3-phosphate - GAPDH glyceraldehyde 3-phosphate dehydrogenase  相似文献   

18.
KLHL3 is a BTB-BACK-Kelch family protein that serves as a substrate adapter in Cullin3 (Cul3) E3 ubiquitin ligase complexes. KLHL3 is highly expressed in distal nephron tubules where it is involved in the regulation of electrolyte homeostasis and blood pressure. Mutations in KLHL3 have been identified in patients with inherited hypertension disorders, and several of the disease-associated mutations are located in the presumed Cul3 binding region. Here, we report the crystal structure of a complex between the KLHL3 BTB-BACK domain dimer and two copies of an N terminal fragment of Cul3. We use isothermal titration calorimetry to directly demonstrate that several of the disease mutations in the KLHL3 BTB-BACK domains disrupt the association with Cul3. Both the BTB and BACK domains contribute to the Cul3 interaction surface, and an extended model of the dimeric CRL3 complex places the two E2 binding sites in a suprafacial arrangement with respect to the presumed substrate-binding sites.  相似文献   

19.
1-Phosphofructokinase (EC 2.7.1.56) (1PFK) was purified and characterized for the first time from an archaebacterial halophile Haloarcula vallismortis. The purification procedure involving (NH4)2SO4 fractionation, (NH4)2SO4-mediated chromatography on Sepharose 4B, CM-cellulose chromatography, hydrophobic on phenyl Sepharose and adsorption chromatography on hydroxylapatite yielded a preparation with a specific activity of 128 and 100-fold purification. From gel filtration and sucrose density gradient ultracentrifugation, the apparent molecular mass of halobacterial 1PFK was found as 76 ± 5 kDa. The halobacterial 1PFK appears to be monomeric and the possibility of an unstable phosphoenzyme intermediate during its catalysis could not be ruled out. As in the case of many halobacterial enzymes, the 1PFK was found to be halophilic and thermostable. Other catalytic features of halobacterial 1PFK were similar to its counterparts from eubacterial sources.  相似文献   

20.
A pre-requisite for bacterial pathogenesis is the successful interaction of a pathogen with a host. One mechanism used by a broad range of Gram negative bacterial pathogens is to deliver effector proteins directly into host cells through a dedicated type III secretion system where they modulate host cell function. The cycle inhibiting factor (Cif) family of effector proteins, identified in a growing number of pathogens that harbour functional type III secretion systems and have a wide host range, arrest the eukaryotic cell cycle. Here, the crystal structures of Cifs from the insect pathogen/nematode symbiont Photorhabdus luminescens (a γ-proteobacterium) and human pathogen Burkholderia pseudomallei (a β-proteobacterium) are presented. Both of these proteins adopt an overall fold similar to the papain sub-family of cysteine proteases, as originally identified in the structure of a truncated form of Cif from Enteropathogenic E. coli (EPEC), despite sharing only limited sequence identity. The structure of an N-terminal region, referred to here as the ‘tail-domain’ (absent in the EPEC Cif structure), suggests a surface likely to be involved in host-cell substrate recognition. The conformation of the Cys-His-Gln catalytic triad is retained, and the essential cysteine is exposed to solvent and addressable by small molecule reagents. These structures and biochemical work contribute to the rapidly expanding literature on Cifs, and direct further studies to better understand the molecular details of the activity of these proteins.  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