首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
Computational small molecule docking into comparative models of proteins is widely used to query protein function and in the development of small molecule therapeutics. We benchmark RosettaLigand docking into comparative models for nine proteins built during CASP8 that contain ligands. We supplement the study with 21 additional protein/ligand complexes to cover a wider space of chemotypes. During a full docking run in 21 of the 30 cases, RosettaLigand successfully found a native-like binding mode among the top ten scoring binding modes. From the benchmark cases we find that careful template selection based on ligand occupancy provides the best chance of success while overall sequence identity between template and target do not appear to improve results. We also find that binding energy normalized by atom number is often less than −0.4 in native-like binding modes.  相似文献   

3.
4.
To establish a proteomic reference map for soybean leaves, we separated and identified leaf proteins using two-dimensional polyacrylamide gel electrophoresis (2D-PAGE) and mass spectrometry (MS). Tryptic digests of 260 spots were subjected to peptide mass fingerprinting (PMF) by matrix-assisted laser desorption/ionization-time of flight (MALDI-TOF) MS. Fifty-three of these protein spots were identified by searching NCBInr and SwissProt databases using the Mascot search engine. Sixty-seven spots that were not identified by MALDI-TOF-MS analysis were analyzed with liquid chromatography tandem mass spectrometry (LC-MS/MS), and 66 of these spots were identified by searching against the NCBInr, SwissProt and expressed sequence tag (EST) databases. We have identified a total of 71 unique proteins. The majority of the identified leaf proteins are involved in energy metabolism. The results indicate that 2D-PAGE, combined with MALDI-TOF-MS and LC-MS/MS, is a sensitive and powerful technique for separation and identification of soybean leaf proteins. A summary of the identified proteins and their putative functions is discussed.  相似文献   

5.
Prp43p catalyzes essential steps in pre-mRNA splicing and rRNA biogenesis. In splicing, Spp382p stimulates the Prp43p helicase to dissociate the postcatalytic spliceosome and, in some way, to maintain the integrity of the spliceosome assembly. Here we present a dosage interference assay to identify Spp382p-interacting factors by screening for genes that when overexpressed specifically inhibit the growth of a conditional lethal prp38-1 spliceosome assembly mutant in the spp382-1 suppressor background. Identified, among others, are genes encoding the established splicing factors Prp8p, Prp9p, Prp11p, Prp39p, and Yhc1p and two poorly characterized proteins with possible links to splicing, Sqs1p and Cwc23p. Sqs1p copurifies with Prp43p and is shown to bind Prp43p and Spp382p in the two-hybrid assay. Overexpression of Sqs1p blocks pre-mRNA splicing and inhibits Prp43p-dependent steps in rRNA processing. Increased Prp43p levels buffer Sqs1p cytotoxicity, providing strong evidence that the Prp43p DExD/H-box protein is a target of Sqs1p. Cwc23p is the only known yeast splicing factor with a DnaJ motif characteristic of Hsp40-like chaperones. We show that similar to SPP382, CWC23 activity is critical for efficient pre-mRNA splicing and intron metabolism yet, surprisingly, this activity does not require the canonical DnaJ/Hsp40 motif. These and related data establish the value of this dosage interference assay for finding genes that alter cellular splicing and define Sqs1p and Cwc23p as prospective modulators of Spp382p-stimuated Prp43p function.EIGHT phylogenetically conserved DExD/H-box proteins act at discrete steps to regulate the assembly, activation, and dissociation of the splicing apparatus (reviewed in Brow 2002; Konarska and Query 2005; Linder 2006). How these RNA-dependent ATPases are temporally and functionally regulated remains poorly understood. One putative regulator, the 83-kDa Spp382/Ntr1 protein (henceforth referred to by the Saccharomyces Genome Database standard name, Spp382p), was discovered in a screen for mutants capable of suppressing defects in yeast spliceosome assembly (Pandit et al. 2006). While spp382 null alleles are lethal, partial loss of function suppresses mutations in several other splicing factors, including the genes encoding the essential spliceosomal proteins Prp8p and Prp38p. Spp382p is a spliceosomal protein that binds the Prp43p DExD/H-box protein to promote efficient dissociation of spliceosomal factors after completion of splicing in vitro (Tsai et al. 2005). Some but not all spp382 mutants also accumulate the excised intron product of splicing in vivo, ostensibly due to protection of the intron within a hyperstabilized spliceosome (Pandit et al. 2006; Tanaka et al. 2007). The suppression of spliceosome assembly defects by spp382 mutation is proposed to occur by impairing Spp382p-stimulated dissociation of kinetically impaired or otherwise inefficient spliceosomes by Prp43p (Pandit et al. 2006). Consistent with this hypothesis, prp43 mutations also suppress spliceosome assembly defects in a manner that, within limits, is inversely proportional to the residual ATPase activity of Prp43p (Pandit et al. 2006). In this light, it is possible that the Spp382 and Prp43 proteins are components of the “discard pathway” for spliceosome dissociation predicted by the kinetic proofreading model of DExD/H-box protein function (Burgess et al. 1990; Konarska and Query 2005).Recently, several groups made the surprising observation that the Prp43p splicing factor is also required for ribosome biogenesis. Mutations in PRP43 inhibit 35S pre-rRNA cleavage and limit downstream steps in this processing pathway (Lebaron et al. 2005; Combs et al. 2006; Leeds et al. 2006). Prp43p is ≥10-fold more abundant than the splicing-restricted DExD/H-box proteins (e.g., Brr2p, Prp2p, Prp5, Prp16, Prp22p, and Prp28p; see Ghaemmaghami et al. 2003) and, consistent with dual function in splicing and rRNA processing, nuclear Prp43p is enriched in the nucleolus (Huh et al. 2003). Furthermore, proteins and small RNAs acting exclusively in splicing or in rRNA biogenesis copurify with Prp43p, supporting its direct contribution to both processes (Ho et al. 2002; Lebaron et al. 2005; Combs et al. 2006; Gavin et al. 2006; Krogan et al. 2006; Leeds et al. 2006). While it is not certain how Prp43p is partitioned within the cell, its association with Spp382p appears critical for recruitment to the postcatalytic spliceosome and for stimulation of the intrinsic Prp43p helicase activity (Tsai et al. 2005; Boon et al. 2006; Pandit et al. 2006; Tanaka et al. 2007). A structurally related protein, Pxr1p, interacts with Prp43p (Lebaron et al. 2005) and is required for efficient rRNA processing (Guglielmi and Werner 2002). Pxr1p may serve a parallel role for Prp43p recruitment and activation within the rRNA processing apparatus but direct evidence for such function is lacking.Here we describe a genetic approach to identify factors that interact with SPP382 and function in spliceosome dynamics. Specifically, we describe a dosage interference assay to find genes that when overexpressed inhibit the growth of a yeast strain in which the temperature-sensitive prp38-1 spliceosome assembly mutation (Xie et al. 1998) is suppressed by the spp382-1 suppressor allele (Pandit et al. 2006). We identify multiple GAL1-effector genes that preferentially inhibit growth of the prp38-1 spp382-1 double mutant compared with either single-mutant host or a wild-type yeast strain. Consistent with the goal of this screen, most genes cause splicing inhibition with galactose induction.Among the recovered genes are SQS1 and CWC23, which encode proteins that purify from yeast in multisubunit protein complexes containing Prp43p but have unknown functions (Lebaron et al. 2005; Pandit et al. 2006). Sqs1p is a nonessential 87-kDa protein that, similar to the Prp43p interacting proteins Pxr1p and Spp382p, possesses the glycine-rich G-patch motif common to a subset of RNA binding proteins (Aravind and Koonin 1999). CWC23 encodes a 33-kDa protein with a canonical DnaJ motif characteristic of Hsp40-like activators of Hsp70 chaperones (Walsh et al. 2004; Vos et al. 2008). Cwc23p interacts with Spp382p in the two-hybrid assay and by affinity selection (Pandit et al. 2006). The genetic and biochemical results presented here establish Sqs1p and Cwc23p as Spp382p-interacting proteins with contributions to RNA processing distinct from what might be expected on the basis of their protein motif characteristics.  相似文献   

6.
By the use of the proteolytic substrates benzoyl-dl-arginine-p-nitroanilide and benzoyl-l-arginine ethyl ester the enzyme arachain has been purified 325-fold from acetone powders of ungerminated peanuts. The pH optimum for the hydrolysis of benzoyl-dl-arginine-p-nitroanilide was 8.1 in tris buffer, and for benzoyl-l-arginine ethyl ester was 7.5 using N - 2 - hydroxyethylpiperazine - N′ - 2 - ethanesulfonic acid buffer. The purest fraction showed one main band with one to three minor bands on disc gel electrophoresis. The major protein component had an S20,w of 6.20. The energy of activation for the hydrolysis of benzoyl-dl-arginine-p-nitroanilide was calculated to be 16 kilocalories. The Michaelis constant for benzoyl-dl-arginine-p-nitroanilide was 10 micromolar and for benzoyl-l-arginine ethyl ester was 110 micromolar. The enzyme showed essentially no activity with casein, dimethyl casein, or bovine serum albumin as substrates. A large number of peptides were hydrolyzed by the enzyme, only l-leucyl-l-tyrosine being resistant of the peptides tested. The results suggest that arachain is not a “trypsin-like” protease but is a peptide hydrolase.  相似文献   

7.
We have developed a novel automated system to analyze protein complexes by integrating a surface plasmon resonance (SPR) biosensor with highly sensitive nanoflow liquid chromatography-tandem mass spectrometry (LC-MS/MS). A His-tagged protein, which is also tagged with FLAG and biotinylated sequences, was expressed in mammalian cells. After purification by using the His tag from the cell lysate, the sample protein mixture was applied to an SPR biosensor and the protein complex was captured on the sensor chip. The automated SPR-LC-MS/MS was then performed: (1) two-step on-chip purification of the protein complex by using the FLAG and the biotinylated tags, (2) on-chip protease digestion of the complex, and (3) online nanoflow LC-MS/MS analysis of the resulting peptide fragments for protein identification. All of these processes could be monitored in real-time by the SPR biosensor. We validated the performance of the system using either FK506-binding protein 52 kDa (FKBP52) or ribosomal protein S19 (rpS19) as bait. Thus, the fully automated SPR-LC-MS/MS system appeared to be a powerful tool for functional proteomics studies, particularly for snapshot analysis of functional cellular complexes and machines.  相似文献   

8.
In Duchenne muscular dystrophy (DMD) patients and the mdx mouse model of DMD, chronic activation of the classical nuclear factor-κB (NF-κB) pathway contributes to the pathogenesis that causes degeneration of muscle fibers, inflammation and fibrosis. Prior studies demonstrate that inhibition of inhibitor of κB kinase (IKK)-mediated NF-κB activation using l-isomer NF-κB essential modulator (NEMO)-binding domain (NBD) peptide-based approaches reduce muscle pathology in the mdx mouse. For our studies, the NBD peptide is synthesized as a fusion peptide with an eight-lysine (8K) protein transduction domain to facilitate intracellular delivery. We hypothesized that the d-isoform peptide could have a greater effect than the naturally occurring l-isoform peptide due to the longer persistence of the d-isoform peptide in vivo. In this study, we compared systemic treatment with low (1 mg/kg) and high (10 mg/kg) doses of l- and d-isomer 8K-wild-type-NBD peptide in mdx mice. Treatment with both l- or d-isoform 8K-wild-type-NBD peptide resulted in decreased activation of NF-κB and improved histology in skeletal muscle of the mdx mouse. However, we observed kidney toxicity (characterized by proteinuria), increased serum creatinine, activation of NF-κB and pathological changes in kidney cortex that were most severe with treatment with the d-isoform of 8K-wild-type-NBD peptide. The observed toxicity was also seen in normal mice.  相似文献   

9.
1. The effects of alkylating agents and disulphides on the thiol-containing proteins of nuclei from rat thymus and liver were studied. Three protein fractions were examined: histones extracted with 50mm- and 250mm-hydrochloric acid and the residual protein. None of the reagents selectively reacted with any one of the protein fractions. 2. Amino acid uptake in vitro into the histones of nuclei from rat thymus was analysed by preparative electrophoresis of the proteins extracted with 50mm- and 250mm-hydrochloric acid. After 1hr. at 37° the greater incorporation was into the proteins extracted with 50mm-hydrochloric acid. 3. Preparative electrophoresis was used to study the relative thiol contents of the proteins of the 50mm-hydrochloric acid extract from thymus nuclei by labelling the histones in vitro with 14C-labelled N-ethylmaleimide. 4. The capacity of the proteins extracted from rat thymus with 50mm- and 250mm-hydrochloric acid, and of the components from these extracts separated by preparative electrophoresis, to combine with DNA and to depress DNA-dependent RNA synthesis was studied. The histones extracted with 50mm-hydrochloric acid were more lysine-rich than those extracted with 250mm-hydrochloric acid. Wide variations were found in the abilities of the separated components to depress RNA synthesis.  相似文献   

10.
11.
Electrometric and spectrophotometric titrations showed vancomycin to contain groups having pK values of about 2.9, 7.2, 8.6, 9.6, 10.5 and 11.7. Of these the four last-named were phenolic. Titration above pH11 and below pH1 was irreversible and antibiotic potency was destroyed. Combination with the specific peptide diacetyl-l-lysyl-d-alanyl-d-alanine hindered the titration of the first three phenolic groups. Spectrophotometric titration of iodovancomycin showed that the phenolic group with pK 9.6 was the one iodinated. The stability of the vancomycin–peptide complex in the range pH1–13 showed that complex-formation occurred only when carboxyl groups were ionized and the phenolic groups were non-ionized. The complex was formed in concentrations of urea up to 8m, of potassium chloride up to 4m, of sodium dodecyl sulphate up to 1%, and at temperatures up to 60°C. From titration curves, organic chlorine and iodine analysis, and combination with peptide, a minimum molecular weight for vancomycin of 1700–1800 was estimated. Optical-rotatory-dispersion and circular-dichroism experiments suggested that vancomycin has only limited conformational flexibility. Both vancomycin and its complexes with peptide exhibited properties suggesting aggregation. Vancomycin and iodovancomycin can be fractionated into a main fraction and at least three minor components. The isolation of these fractions salt-free is described and their antibiotic properties are shown to correlate with their ability to form complexes with peptide.  相似文献   

12.
13.
Evidence for a specific glutamate/h cotransport in isolated mesophyll cells   总被引:1,自引:1,他引:0  
Mechanically isolated Asparagus sprengeri Regel mesophyll cells were suspended in 1 millimolar CaSO4. Immediate alkalinization of the medium occured on the addition of 1 millimolar concentrations of l-glutamate (Glu) and its analog l-methionine-d,l-sulfoximine (l-MSO). d-Glu and the l isomers of the protein amino acids did not elicit alkalinization. l-Glu dependent alkalinization was transient and acidification resumed after approximately 30 to 45 minutes. At pH 6.0, 5 millimolar l-Glu stimulated initial rates of alkalinization that varied between 1.3 to 4.1 nmol H+/106 cells·minute. l-Glu dependent alkalinization was saturable, increased with decreasing pH, was inhibited by carbonyl cyanide-p-trichloromethoxyphenyl hydrazone (CCCP), and was not stimulated by light. Uptake of l-[U-14C]glutamate increased as the pH decreased from 6.5 to 5.5, and was inhibited by l-MSO. l-Glu had no influence on K+ efflux. Although evidence for multiple amino acid/proton cotransport systems has been found in other tissues, the present report indicates that a highly specific l-Glu/proton uptake process is present in Asparagus mesophyll cells.  相似文献   

14.
Comparative genomics provides a powerful tool for the identification of genes that encode traits shared between crop plants and model organisms. Pathogen resistance conferred by plant R genes of the nucleotide-binding–leucine-rich-repeat (NB–LRR) class is one such trait with great agricultural importance that occupies a critical position in understanding fundamental processes of pathogen detection and coevolution. The proposed rapid rearrangement of R genes in genome evolution would make comparative approaches tenuous. Here, we test the hypothesis that orthology is predictive of R-gene genomic location in the Solanaceae using the pepper R gene Bs2. Homologs of Bs2 were compared in terms of sequence and gene and protein architecture. Comparative mapping demonstrated that Bs2 shared macrosynteny with R genes that best fit criteria determined to be its orthologs. Analysis of the genomic sequence encompassing solanaceous R genes revealed the magnitude of transposon insertions and local duplications that resulted in the expansion of the Bs2 intron to 27 kb and the frequently detected duplications of the 5′-end of R genes. However, these duplications did not impact protein expression or function in transient assays. Taken together, our results support a conservation of synteny for NB–LRR genes and further show that their distribution in the genome has been consistent with global rearrangements.R genes have a central role in plant disease resistance to mediate pathogen detection and response (Martin et al. 2003; Glazebrook 2005). Although R genes are only one of the components required for these responses, they are consistently identified as a critical determinant for qualitative and quantitative resistance (Fluhr 2001; Wisser et al. 2006). The structure, mechanism of action, and evolution of this gene family are still being elucidated and are critical issues for a more efficient deployment of disease resistances in agricultural crops (McDowell and Simon 2006; Takken et al. 2006; Friedman and Baker 2007; van Ooijen et al. 2007).Comparative studies of sequence similarity between plant R proteins and proteins of innate immunity in animals have made important contributions toward understanding R-protein structure, the role of individual protein domains, and the mechanism by which R proteins identify and respond to foreign proteins (Nurnberger et al. 2004; Takken et al. 2006; Rairdan and Moffett 2007). Both share a central nucleotide-binding (NB) site and a region of homology termed the “ARC” domain (collectively referred to as the NB–ARC) (van der Biezen and Jones 1998; Rairdan and Moffett 2007). The plant counterparts have a highly variable leucine-rich-repeat (LRR) domain at the C terminus and, at the N terminus, either a domain with homology to the Toll and interleukin-1 receptors (TIR) or lack this feature, instead possessing a domain that may include a coiled-coil motif. Due to uncertainty regarding the presence of a coiled-coil motif, this class of NB–LRRs is often referred to as non-TIR proteins. The LRR domains are highly variable and tend to be under diversifying selection to adapt to continually changing pathogen proteins (Meyers et al. 1998b; Michelmore and Meyers 1998; Mondragon-Palomino et al. 2002). Other conserved patterns have been identified in the N terminus of non-TIR proteins, most notably, an EDxxD motif that mediates an intramolecular interaction (Rairdan et al. 2008). The interaction with cellular factors is mediated by the N-terminal domains of NB–LRR proteins although domain-swapping experiments between closely related NB–LRR proteins have shown that recognition specificity is determined by the LRR domains (Rairdan and Moffett 2007; van Ooijen et al. 2007).The clustering of R genes has provided both insight into their ability to evolve rapidly and challenges to their identification and cloning. R genes often occur in clusters of tandem duplications that can span several megabases and include a multitude of copies of functional R genes, pseudogenes, and other genes within the clusters (Meyers et al. 1998a; Kuang et al. 2004; Smith et al. 2004). Of the various modes of evolution ascribed to these clusters, sequence exchange between R genes within the cluster by unequal crossing over or illegitimate recombination is especially noteworthy (Michelmore and Meyers 1998; Ellis et al. 2000; Hulbert et al. 2001; McDowell and Simon 2006; Friedman and Baker 2007; Wicker et al. 2007). Under stress conditions, transposon activation, recombination activation, and chromatin modifications related to small RNAs may be induced (Levy et al. 2004; Friedman and Baker 2007; Yi and Richards 2007).Two distinct models for the genomewide arrangement and distribution of NB–LRR genes and these clusters have been proposed. The first predicts rapid rearrangement of R-gene distribution during genome evolution, yielding poor conservation of R-gene locations (Leister et al. 1998; Richly et al. 2002; Meyers et al. 2003). Indeed, in monocots, extensive loss of genomewide R-gene colinearity has been attributed to frequent R-gene duplication and ectopic transposition (Gale and Devos 1998; Paterson et al. 2003). In contrast, the second model supports genomewide conservation of R-gene distribution maintained during speciation. According to this model, most duplication and recombination of R-gene sequences should occur within restricted chromosomal regions, yielding clusters of closely related R-gene sequences. The resulting orthology relationships (homologs related by speciation, not duplication) are complex due to “fractionation” (repeated cycles of duplication, deletion, and recombination) but can, as we have previously shown, be reconstructed (Grube et al. 2000b). Analysis of R genes using the complete Arabidopsis thaliana genome sequence supports this model and accounts for the consensus of NB–LRR sequences (Baumgarten et al. 2003). Resistance to a particular pathogen type is not conserved, and highly similar NB–LRR proteins may confer resistance to very different pathogens (Grube et al. 2000b).Bs2 encodes a non-TIR NB–LRR protein identified in Capsicum chacoense that confers resistance to the bacterium Xanthomonas campestris pv. vesicatoria. This R gene has greatest sequence identity to Rx and Gpa2 in potato, which confer resistance to a virus and nematode, respectively (Bendahmane et al. 1999; Tai et al. 1999b; van der Vossen et al. 2000). Despite the difference in the pathogens recognized by these genes, they are distinguishable from all other known R genes by marked sequence and structural features. In this study, we demonstrate that these three R genes are derived from syntenic regions in solanaceous genomes as predicted by our model of conservation of synteny. In performing these comparisons, we explore conserved amino acid patterns associated with proteins of the non-TIR family and the local genomic context of R genes of the Solanaceae. Finally, advances in the development of the Solanaceae as a system for comparative genomics highlight a role for chromosomal rearrangements in R-gene distribution throughout plant genomes.  相似文献   

15.
Hart JW  Filner P 《Plant physiology》1969,44(9):1253-1259
The sulfur requirements of tobacco (Nicotiana tabacum L. var. Xanthi) XD cells grown in chemically defined liquid media can be satisfied by sulfate, thiosulfate, l-cyst(e)ine, l-methionine or glutathione, and somewhat less effectively by d-cyst (e) ine, d-methionine or dl-homocyst (e)ine. Sulfate uptake is inhibited after a 2 hr lag by l-cyst (e)ine, l-methionine, l-homocyst(e)ine or l-isoleucine, but not by any of the other protein amino acids, nor by d-cyst(e)ine. l-cyst(e)ine is neither a competitive nor a non-competitive inhibitor of sulfate uptake. Its action most closely resembles apparent uncompetitive inhibition. Inhibition of sulfate uptake by l-cyst(e)ine can be partially prevented by equimolar l-arginine, l-lysine, l-leucine, l-phenylalanine, l-tyrosine or l-tryptophan, but is little affected by any of the other protein amino acids. The effective amino acids are apparent competitive inhibitors of l-cyst(e)ine uptake after a 2 hr lag. Inhibition of sulfate uptake by l-methionine cannot be prevented, nor can uptake of l-methionine be inhibited by any single protein amino acid. The results suggest the occurrence of negative feedback control of sulfate assimilation by the end products, the sulfur amino acids, in cultured tobacco cells.  相似文献   

16.
d-tyrosyl-tRNATyr deacylase (DTD) is an editing enzyme that removes d-amino acids from mischarged tRNAs. We describe an in-depth analysis of the malaria parasite Plasmodium falciparum DTD here. Our data provide structural insights into DTD complexes with adenosine and d-amino acids. Bound adenosine is proximal to the DTD catalysis site, and it represents the authentic terminal adenosine of charged tRNA. DTD-bound d-amino acids cluster at three different subsites within the overall active site pocket. These subsites, called transition, active, and exit subsites allow docking, re-orientation, chiral selection, catalysis, and exit of the free d-amino acid from DTD. Our studies reveal variable modes of d-amino acid recognition by DTDs, suggesting an inherent plasticity that can accommodate all d- amino acids. An in-depth analysis of native, ADP-bound, and d- amino acid-complexed DTD structures provide the first atomic snapshots of ligand recognition and subsequent catalysis by this enzyme family. We have mapped sites for the deacylation reaction and mark possible routes for entry and egress of all substrates and products. We have also performed structure-based inhibitor discovery and tested lead compounds against the malaria parasite P. falciparum using growth inhibition assays. Our studies provide a comprehensive structural basis for the catalytic mechanism of DTD enzymes and have implications for inhibition of this enzyme in P. falciparum as a route to inhibiting the parasite.  相似文献   

17.
Protein kinases are important mediators of signal transduction in eukaryotic cells, and identifying the substrates of these enzymes is essential for a complete understanding of most signaling networks. In this report, novel substrate-binding variants of the cAMP-dependent protein kinase (PKA) were used to identify substrate domains required for efficient phosphorylation in vivo. Most wild-type protein kinases, including PKA, interact only transiently with their substrates. The substrate domains identified were distal to the sites of phosphorylation and were found to interact with a C-terminal region of PKA that was itself removed from the active site. Only a small set of PKA alterations resulted in a stable association with substrates, and the identified residues were clustered together within the hydrophobic core of this enzyme. Interestingly, these residues stretched from the active site of the enzyme to the C-terminal substrate-binding domain identified here. This spatial organization is conserved among the entire eukaryotic protein kinase family, and alteration of these residues in a second, unrelated protein kinase also resulted in a stable association with substrates. In all, this study identified distal sites in PKA substrates that are important for recognition by this enzyme and suggests that the interaction of these domains with PKA might influence specific aspects of substrate binding and/or release.PROTEIN kinases are key mediators of signal transduction in all eukaryotic cells. Each protein kinase modifies a distinct set of substrates, and the biological consequences of activating any kinase are the result of the collective actions of these target proteins (Hunter 2000; Manning et al. 2002). The ability to identify substrates is therefore essential for a complete understanding of most signaling pathways. Unfortunately, this identification process tends to be difficult, and few physiologically relevant targets are known for most protein kinases (Manning and Cantley 2002; Johnson and Hunter 2005). This situation may be changing as a number of innovative approaches to this problem have been developed in recent years (reviewed in Ptacek and Snyder 2006; Deminoff and Herman 2007; Ubersax and Ferrell 2007).This article is focused on the cAMP-dependent protein kinase (PKA) from the budding yeast, Saccharomyces cerevisiae. The PKA enzyme is found in all eukaryotes and is one of the most intensely studied members of this protein family (Taylor et al. 2005). PKA was the first protein kinase structure to be described, and its structure has provided essential insights into the general organization and catalytic mechanism of these enzymes (Knighton et al. 1991; Smith et al. 1999). Subsequent work has illustrated the conserved nature of the protein kinase core and the different ways that the activity of these enzymes can be regulated (Hunter 2000; Huse and Kuriyan 2002; Kannan and Neuwald 2005). In S. cerevisiae, PKA activity is a key regulator of cell growth and the response to environmental stress (Toda et al. 1985; Thevelein and De Winde 1999; Herman 2002; Schneper et al. 2004). We are interested in understanding the role of PKA in these processes and have identified a number of substrates for this enzyme (Howard et al. 2003; Chang et al. 2004; Budovskaya et al. 2005; Deminoff et al. 2006). One of the approaches used for this identification took advantage of PKA variants that exhibit a stable binding to substrate proteins (Deminoff et al. 2006). This binding is novel as most wild-type protein kinases, including PKA, interact only transiently with their substrates (Manning and Cantley 2002). Interestingly, one of these PKA variants was altered at a residue that is conserved in all protein kinases, suggesting that it might be possible to generate substrate-binding versions of other enzymes in this family.These variants of PKA were used here to explore the nature of the protein kinase–substrate interaction. These studies identified substrate domains distal to the sites of phosphorylation that were required for efficient recognition by the wild-type PKA, both in vitro and in vivo. These substrate domains were found to interact with a C-terminal region of PKA that is itself removed from the active site of the enzyme. A systematic mutagenesis of PKA identified additional residues that, when altered, resulted in a stable association with substrates. These latter residues are in close proximity in the three-dimensional structure and may link the active site with this C-terminal substrate-binding domain of PKA. Finally, we show that similar alterations within a second protein kinase, the mammalian double-stranded RNA-dependent protein kinase (PKR), also led to an increased affinity for substrates. In all, the data suggest that the interactions described here may be generally important for protein kinase function and models that explain potential roles for these substrate domains are discussed.  相似文献   

18.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

19.
The soil bacterium Bacillus subtilis forms biofilms on surfaces and at air-liquid interfaces. It was previously reported that these biofilms disassemble late in their life cycle and that conditioned medium from late-stage biofilms inhibits biofilm formation. Such medium contained a mixture of d-leucine, d-methionine, d-tryptophan, and d-tyrosine and was reported to inhibit biofilm formation via the incorporation of these d-amino acids into the cell wall. Here, we show that l-amino acids were able to specifically reverse the inhibitory effects of their cognate d-amino acids. We also show that d-amino acids inhibited growth and the expression of biofilm matrix genes at concentrations that inhibit biofilm formation. Finally, we report that the strain routinely used to study biofilm formation has a mutation in the gene (dtd) encoding d-tyrosyl-tRNA deacylase, an enzyme that prevents the misincorporation of d-amino acids into protein in B. subtilis. When we repaired the dtd gene, B. subtilis became resistant to the biofilm-inhibitory effects of d-amino acids without losing the ability to incorporate at least one noncanonical d-amino acid, d-tryptophan, into the peptidoglycan peptide side chain. We conclude that the susceptibility of B. subtilis to the biofilm-inhibitory effects of d-amino acids is largely, if not entirely, due to their toxic effects on protein synthesis.  相似文献   

20.
The spontaneous degradation of asparaginyl and aspartyl residues to isoaspartyl residues is a common type of protein damage in aging organisms. Although the protein-l-isoaspartyl (d-aspartyl) O-methyltransferase (EC 2.1.1.77) can initiate the repair of l-isoaspartyl residues to l-aspartyl residues in most organisms, no gene homolog or enzymatic activity is present in the budding yeast Saccharomyces cerevisiae. Therefore, we used biochemical approaches to elucidate how proteins containing isoaspartyl residues are metabolized in this organism. Surprisingly, the level of isoaspartyl residues in yeast proteins (50–300 pmol of isoaspartyl residues/mg of protein extract) is comparable with organisms with protein-l-isoaspartyl (d-aspartyl) O-methyltransferase, suggesting a novel regulatory pathway. Interfering with common protein quality control mechanisms by mutating and inhibiting the proteasomal and autophagic pathways in vivo did not increase isoaspartyl residue levels compared with wild type or uninhibited cells. However, the inhibition of metalloproteases in in vitro aging experiments by EDTA resulted in an ∼3-fold increase in the level of isoaspartyl-containing peptides. Characterization by mass spectrometry of these peptides identified several proteins involved in metabolism as targets of isoaspartyl damage. Further analysis of these peptides revealed that many have an N-terminal isoaspartyl site and originate from proteins with short half-lives. These results suggest that one or more metalloproteases participate in limiting isoaspartyl formation by robust proteolysis.  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