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 15 毫秒
1.
The N‐end rule pathway is conserved from bacteria to man and determines the half‐life of a protein based on its N‐terminal amino acid. In Escherichia coli, model substrates bearing an N‐degron are recognised by ClpS and degraded by ClpAP in an ATP‐dependent manner. Here, we report the isolation of 23 ClpS‐interacting proteins from E. coli. Our data show that at least one of these interacting proteins—putrescine aminotransferase (PATase)—is post‐translationally modified to generate a primary N‐degron. Remarkably, the N‐terminal modification of PATase is generated by a new specificity of leucyl/phenylalanyl‐tRNA‐protein transferase (LFTR), in which various combinations of primary destabilising residues (Leu and Phe) are attached to the N‐terminal Met. This modification (of PATase), by LFTR, is essential not only for its recognition by ClpS, but also determines the stability of the protein in vivo. Thus, the N‐end rule pathway, through the ClpAPS‐mediated turnover of PATase may have an important function in putrescine homeostasis. In addition, we have identified a new element within the N‐degron, which is required for substrate delivery to ClpA.  相似文献   

2.
The N‐end rule pathway uses an evolutionarily conserved mechanism in bacteria and eukaryotes that marks proteins for degradation by ATP‐dependent chaperones and proteases such as the Clp chaperones and proteases. Specific N‐terminal amino acids (N‐degrons) are sufficient to target substrates for degradation. In bacteria, the ClpS adaptor binds and delivers N‐end rule substrates for their degradation upon association with the ClpA/P chaperone/protease. Here, we report the first crystal structure, solved at 2.7 Å resolution, of a eukaryotic homolog of bacterial ClpS from the malaria apicomplexan parasite Plasmodium falciparum (Pfal). Despite limited sequence identity, Plasmodium ClpS is very similar to bacterial ClpS. Akin to its bacterial orthologs, plasmodial ClpS harbors a preformed hydrophobic pocket whose geometry and chemical properties are compatible with the binding of N‐degrons. However, while the N‐degron binding pocket in bacterial ClpS structures is open and accessible, the corresponding pocket in Plasmodium ClpS is occluded by a conserved surface loop that acts as a latch. Despite the closed conformation observed in the crystal, we show that, in solution, Pfal‐ClpS binds and discriminates peptides mimicking bona fide N‐end rule substrates. The presence of an apicoplast targeting peptide suggests that Pfal‐ClpS localizes to this plastid‐like organelle characteristic of all Apicomplexa and hosting most of its Clp machinery. By analogy with the related ClpS1 from plant chloroplasts and cyanobacteria, Plasmodium ClpS likely functions in association with ClpC in the apicoplast. Our findings open new venues for the design of novel anti‐malarial drugs aimed at disrupting parasite‐specific protein quality control pathways.  相似文献   

3.
4.
The N‐end rule relates the in vivo half‐life of a protein to the identity of its N‐terminal amino acid residue. While some N‐terminal residues result in metabolically stable proteins, other, so‐called destabilizing residues, lead to rapid protein turnover. The N‐end rule pathway, which mediates the recognition and degradation of proteins with N‐terminal destabilizing residues, is present in all organisms examined, including prokaryotes. This protein degradation pathway has a hierarchical organization in which some N‐terminal residues, called primary destabilizing residues, are directly recognized by specific ubiquitin ligases. Other destabilizing residues, termed secondary and tertiary destabilizing residues, require modifications before the corresponding proteins can be targeted for degradation by ubiquitin ligases. In eukaryotes, the N‐end rule pathway is a part of the ubiquitin/proteasome system and is known to play essential roles in a broad range of biological processes in fungi, animals and plants. While the structure of the N‐end rule pathway has been extensively studied in yeast and mammals, knowledge of its organization in plants is limited. Using both tobacco and Arabidopsis, we identified the complete sets destabilizing and stabilizing N‐terminal residues. We also characterized the hierarchical organization of the plant N‐end rule by identifying and determining the specificity of two distinct N‐terminal amidohydrolases (Nt‐amidases) of Arabidopsis that are essential for the destabilizing activity of the tertiary destabilizing residues Asn and Gln. Our results indicate that both the N‐end rule itself and mechanistic aspects of the N‐end rule pathway in angiosperms are very similar to those of mammals.  相似文献   

5.
The N‐degron pathway determines the half‐life of proteins in both prokaryotes and eukaryotes by precisely recognizing the N‐terminal residue (N‐degron) of substrates. ClpS proteins from bacteria bind to substrates containing hydrophobic N‐degrons (Leu, Phe, Tyr, and Trp) and deliver them to the caseinolytic protease system ClpAP. This mechanism is preserved in organelles such as mitochondria and chloroplasts. Bacterial ClpS adaptors bind preferentially to Leu and Phe N‐degrons; however, ClpS1 from Arabidopsis thaliana (AtClpS1) shows a difference in that it binds strongly to Phe and Trp N‐degrons and only weakly to Leu. This difference in behavior cannot be explained without structural information due to the high sequence homology between bacterial and plant ClpS proteins. Here, we report the structure of AtClpS1 at 2.0 Å resolution in the presence of a bound N‐degron. The key determinants for α‐amino group recognition are conserved among all ClpS proteins, but the α3‐helix of eukaryotic AtClpS1 is significantly shortened, and consequently, a loop forming a pocket for the N‐degron is moved slightly outward to enlarge the pocket. In addition, amino acid replacement from Val to Ala causes a reduction in hydrophobic interactions with Leu N‐degron. A combination of the fine‐tuned hydrophobic residues in the pocket and the basic gatekeeper at the entrance of the pocket controls the N‐degron selectivity of the plant ClpS protein.  相似文献   

6.
The N‐end rule denotes the relationship between the identity of the amino‐terminal residue of a protein and its in vivo half‐life. Since its discovery in 1986, the N‐end rule has generally been described by a defined set of rules for determining whether an amino‐terminal residue is stabilizing or not. However, recent studies are revealing that this N‐end rule (or N‐degron concept) is less straightforward than previously appreciated. For instance, it is unveiled that N‐terminal acetylation of N‐terminal residues may create a degradation signal (Ac‐degron) that promotes the degradation of target proteins. A recent high‐throughput dissection of degrons in yeast proteins amino termini intriguingly suggested that the hydrophobicity of amino‐terminal residues—but not the N‐terminal acetylation status—may be the indispensable feature of amino‐terminal degrons. Herein, these recent advances in N‐terminal acetylation and the complexity of N‐terminal degradation signals in the context of the N‐degron pathway are analyzed.  相似文献   

7.
The N‐end rule relates the regulation of the in vivo half‐life of a protein to the identity of its N‐terminal residue. Degradation signals (degrons) that are targeted by the N‐end rule pathway include a set called N‐degrons. The main determinant of an N‐degron is a destabilizing N‐terminal residue of a protein. In eukaryotes, the N‐end rule pathway is a part of the ubiquitin system and consists of two branches, the Ac/N‐end rule and the Arg/N‐end rule pathways. The Ac/N‐end rule pathway targets proteins containing Nα‐terminally acetylated (Nt‐acetylated) residues. The Arg/N‐end rule pathway recognizes unacetylated N‐terminal residues and involves N‐terminal arginylation. Together, these branches target for degradation a majority of cellular proteins. For example, more than 80% of human proteins are cotranslationally Nt‐acetylated. Thus, most proteins harbor a specific degradation signal, termed AcN‐degron, from the moment of their birth. Specific N‐end rule pathways are also present in prokaryotes and in mitochondria. Enzymes that produce N‐degrons include methionine‐aminopeptidases, caspases, calpains, Nt‐acetylases, Nt‐amidases, arginyl‐transferases, and leucyl‐transferases. Regulated degradation of specific proteins by the N‐end rule pathway mediates a legion of physiological functions, including the sensing of heme, oxygen, and nitric oxide; selective elimination of misfolded proteins; the regulation of DNA repair, segregation, and condensation; the signaling by G proteins; the regulation of peptide import, fat metabolism, viral and bacterial infections, apoptosis, meiosis, spermatogenesis, neurogenesis, and cardiovascular development; and the functioning of adult organs, including the pancreas and the brain. Discovered 25 years ago, this pathway continues to be a fount of biological insights.  相似文献   

8.
Retinal degenerative diseases (RDs) are a group of inherited diseases characterized by the loss of photoreceptor cells. Selective photoreceptor loss can be induced in mice by an intraperitoneal injection of N‐methyl‐N‐nitrosourea (MNU) and, because of its selectivity, this model is widely used to study the mechanism of RDs. Although it is known that calcium‐calpain activation and lipid peroxidation are involved in the initiation of cell death, the precise mechanisms of this process remain unknown. Heat shock protein 70 (HSP70) has been shown to function as a chaperone molecule to protect cells against environmental and physiological stresses. In this study, we investigated the role of HSP70 on photoreceptor cell death in mice. HSP70 induction by valproic acid, a histone deacetylase inhibitor, attenuated the photoreceptor cell death by MNU through inhibition of apoptotic caspase signals. Furthermore, HSP70 itself was rapidly and calpain‐dependently cleaved after MNU treatment. Therefore, HSP70 induction by valproic acid was dually effective against MNU‐induced photoreceptor cell loss as a result of its anti‐apoptotic actions and its ability to prevent HSP70 degradation. These findings might help lead us to a better understanding of the pathogenic mechanism of RDs.

  相似文献   


9.
Despite extensive understanding of sleep regulation, the molecular‐level cause and function of sleep are unknown. I suggest that they originate in individual neurons and stem from increased production of protein fragments during wakefulness. These fragments are transient parts of protein complexes in which the fragments were generated. Neuronal Ca2+ fluxes are higher during wakefulness than during sleep. Subunits of transmembrane channels and other proteins are cleaved by Ca2+‐activated calpains and by other nonprocessive proteases, including caspases and secretases. In the proposed concept, termed the fragment generation (FG) hypothesis, sleep is a state during which the production of fragments is decreased (owing to lower Ca2+ transients) while fragment‐destroying pathways are upregulated. These changes facilitate the elimination of fragments and the remodeling of protein complexes in which the fragments resided. The FG hypothesis posits that a proteolytic cleavage, which produces two fragments, can have both deleterious effects and fitness‐increasing functions. This (previously not considered) dichotomy can explain both the conservation of cleavage sites in proteins and the evolutionary persistence of sleep, because sleep would counteract deleterious aspects of protein fragments. The FG hypothesis leads to new explanations of sleep phenomena, including a longer sleep after sleep deprivation. Studies in the 1970s showed that ethanol‐induced sleep in mice can be strikingly prolonged by intracerebroventricular injections of either Ca2+ alone or Ca2+ and its ionophore (Erickson et al., Science 1978;199:1219–1221; Harris, Pharmacol Biochem Behav 1979;10:527–534; Erickson et al., Pharmacol Biochem Behav 1980;12:651–656). These results, which were never interpreted in connection to protein fragments or the function of sleep, may be accounted for by the FG hypothesis about molecular causation of sleep.  相似文献   

10.
The expression of intron‐containing genes in eukaryotes requires generation of protein‐coding messenger RNAs (mRNAs) via RNA splicing, whereby the spliceosome removes non‐coding introns from pre‐mRNAs and joins exons. Spliceosomes must ensure accurate removal of highly diverse introns. We show that Sde2 is a ubiquitin‐fold‐containing splicing regulator that supports splicing of selected pre‐mRNAs in an intron‐specific manner in Schizosaccharomyces pombe. Both fission yeast and human Sde2 are translated as inactive precursor proteins harbouring the ubiquitin‐fold domain linked through an invariant GGKGG motif to a C‐terminal domain (referred to as Sde2‐C). Precursor processing after the first di‐glycine motif by the ubiquitin‐specific proteases Ubp5 and Ubp15 generates a short‐lived activated Sde2‐C fragment with an N‐terminal lysine residue, which subsequently gets incorporated into spliceosomes. Absence of Sde2 or defects in Sde2 activation both result in inefficient excision of selected introns from a subset of pre‐mRNAs. Sde2 facilitates spliceosomal association of Cactin/Cay1, with a functional link between Sde2 and Cactin further supported by genetic interactions and pre‐mRNA splicing assays. These findings suggest that ubiquitin‐like processing of Sde2 into a short‐lived activated form may function as a checkpoint to ensure proper splicing of certain pre‐mRNAs in fission yeast.  相似文献   

11.
ArnA from Escherichia coli is a key enzyme involved in the formation of 4‐amino‐4‐deoxy‐l ‐arabinose. The addition of this sugar to the lipid A moiety of the lipopolysaccharide of pathogenic Gram‐negative bacteria allows these organisms to evade the cationic antimicrobial peptides of the host immune system. Indeed, it is thought that such modifications may be responsible for the repeated infections of cystic fibrosis patients with Pseudomonas aeruginosa. ArnA is a bifunctional enzyme with the N‐ and C‐terminal domains catalyzing formylation and oxidative decarboxylation reactions, respectively. The catalytically competent cofactor for the formylation reaction is N10‐formyltetrahydrofolate. Here we describe the structure of the isolated N‐terminal domain of ArnA in complex with its UDP‐sugar substrate and N5‐formyltetrahydrofolate. The model presented herein may prove valuable in the development of new antimicrobial therapeutics.  相似文献   

12.
13.
14.
15.
In mammals and yeast, tail‐anchored (TA) membrane proteins destined for the post‐translational pathway are safely delivered to the endoplasmic reticulum (ER) membrane by a well‐known targeting factor, TRC40/Get3. In contrast, the underlying mechanism for translocation of TA proteins in plants remains obscure. How this unique eukaryotic membrane‐trafficking system correctly distinguishes different subsets of TA proteins destined for various organelles, including mitochondria, chloroplasts and the ER, is a key question of long standing. Here, we present crystal structures of algal ArsA1 (the Get3 homolog) in a distinct nucleotide‐free open state and bound to adenylyl‐imidodiphosphate. This approximately 80‐kDa protein possesses a monomeric architecture, with two ATPase domains in a single polypeptide chain. It is capable of binding chloroplast (TOC34 and TOC159) and mitochondrial (TOM7) TA proteins based on features of its transmembrane domain as well as the regions immediately before and after the transmembrane domain. Several helices located above the TA‐binding groove comprise the interlocking hook‐like motif implicated by mutational analyses in TA substrate recognition. Our data provide insights into the molecular basis of the highly specific selectivity of interactions of algal ArsA1 with the correct sets of TA substrates before membrane targeting in plant cells.  相似文献   

16.
Endo‐βN‐acetylglucosaminidase isolated from B. infantis ATCC 15697 (EndoBI‐1) is a novel enzyme that cleaves N‐N′‐diacetyl chitobiose moieties found in the N‐glycan core of high mannose, hybrid, and complex N‐glycans. These conjugated N‐glycans are recently shown as a new prebiotic source that stimulates the growth of a key infant gut microbe, Bifidobacterium longum subsp. Infantis. The effects of pH (4.45–8.45), temperature (27.5–77.5°C), reaction time (15–475 min), and enzyme/protein ratio (1:3,000–1:333) were evaluated on the release of N‐glycans from bovine colostrum whey by EndoBI‐1. A central composite design was used, including a two‐level factorial design (24) with four center points and eight axial points. In general, low pH values, longer reaction times, higher enzyme/protein ratio, and temperatures around 52°C resulted in the highest yield. The results demonstrated that bovine colostrum whey, considered to be a by/waste product, can be used as a glycan source with a yield of 20 mg N‐glycan/g total protein under optimal conditions for the ranges investigated. Importantly, these processing conditions are suitable to be incorporated into routine dairy processing activities, opening the door for an entirely new class of products (released bioactive glycans and glycan‐free milk). The new enzyme's activity was also compared with a commercially available enzyme, showing that EndoBI‐1 is more active on native proteins than PNGase F and can be efficiently used during pasteurization, streamlining its integration into existing processing strategies. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:1331–1339, 2015  相似文献   

17.
18.
19.
Various aerolysin‐like pore‐forming proteins have been identified from bacteria to vertebrates. However, the mechanism of receptor recognition and/or pore formation of the eukaryotic members remains unknown. Here, we present the first crystal and electron microscopy structures of a vertebrate aerolysin‐like protein from Danio rerio, termed Dln1, before and after pore formation. Each subunit of Dln1 dimer comprises a β‐prism lectin module followed by an aerolysin module. Specific binding of the lectin module toward high‐mannose glycans triggers drastic conformational changes of the aerolysin module in a pH‐dependent manner, ultimately resulting in the formation of a membrane‐bound octameric pore. Structural analyses combined with computational simulations and biochemical assays suggest a pore‐forming process with an activation mechanism distinct from the previously characterized bacterial members. Moreover, Dln1 and its homologs are ubiquitously distributed in bony fishes and lamprey, suggesting a novel fish‐specific defense molecule.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号