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 62 毫秒
1.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

2.
3.
4.
TDP-43 is a nuclear protein involved in exon skipping and alternative splicing. Recently, TDP-43 has been identified as the pathological signature protein in frontotemporal lobar degeneration with ubiquitin-positive inclusions and in amyotrophic lateral sclerosis. In addition, TDP-43-positive inclusions are present in Parkinson disease, dementia with Lewy bodies, and 30% of Alzheimer disease cases. Pathological TDP-43 is redistributed from the nucleus to the cytoplasm, where it accumulates. An ∼25-kDa C-terminal fragment of TDP-43 accumulates in affected brain regions, suggesting that it may be involved in the disease pathogenesis. Here, we show that overexpression of the 25-kDa C-terminal fragment is sufficient to cause the mislocalization and cytoplasmic accumulation of endogenous full-length TDP-43 in two different cell lines, thus recapitulating a key biochemical characteristic of TDP-43 proteinopathies. We also found that TDP-43 mislocalization is associated with a reduction in the low molecular mass neurofilament mRNA levels. Notably, we show that the autophagic system plays a role in TDP-43 metabolism. Specifically, we found that autophagy inhibition increases the accumulation of the C-terminal fragments of TDP-43, whereas inhibition of mTOR, a key protein kinase involved in autophagy regulation, reduces the 25-kDa C-terminal fragment accumulation and restores TDP-43 localization. Our results suggest that autophagy induction may be a valid therapeutic target for TDP-43 proteinopathies.TDP-43 (transactive response DNA-binding protein 43) is a conserved and ubiquitously expressed nuclear protein with a theoretical molecular mass of ∼44 kDa. It is encoded by the TARDBP gene on chromosome 1, which is made of six exons that can be alternatively spliced to yield 11 different isoforms, with the mRNA encoding TDP-43 being the major species (1). Functionally, TDP-43 appears to be involved in exon skipping and alternative splicing (2, 3), and it has also been shown to link different types of nuclear bodies (4). Structural studies have confirmed the presence of two RNA recognition motifs (RRM1 and RRM2) and a glycine-rich C-terminal tail, which is thought to mediate protein-protein interaction (5).Recently, TDP-43 has been shown to be the major pathological protein in a wide range of disorders referred to as TDP-43 proteinopathies (68). These include frontotemporal lobar degeneration with ubiquitin-positive inclusions (FTLD-U),2 motor neuron disease, and amyotrophic lateral sclerosis (ALS). These last two disorders have been directly linked to mutations in TDP-43 (9, 10). In addition, TDP-43-positive inclusions are present in Parkinson disease, dementia with Lewy bodies, and 30% of Alzheimer disease cases (1114). Sporadic and familial forms of FTLD-U and ALS are characterized by cytoplasmic accumulation of insoluble, hyperphosphorylated, ubiquitinated, and proteolytically cleaved C-terminal fragments in affected brain and spinal cord regions. The cytoplasmic accumulation of TDP-43 is associated with a depletion of nuclear TDP-43 (8, 1521). These data suggest that some of these TDP-43 proteinopathies may share common mechanisms of pathogenesis.FTLD-U is caused by loss-of-function mutations in the progranulin gene, which lead, by an unknown mechanism, to the accumulation of cytoplasmic TDP-43 inclusions (22, 23). Notably, the TDP-43 inclusions in the ALS and FTLD-U brains are enriched with TDP-43 C-terminal fragments (8, 19). It has been suggested that the C-terminal fragments can be obtained by caspase-dependent cleavage of the full-length protein (24). However, it remains to be established if these fragments play a role in the disease pathogenesis.TDP-43 proteinopathies are characterized by the accumulation of abnormally modified TDP-43, suggesting that dysfunction in the intracellular quality control systems (ubiquitin-proteasome system and the autophagy-lysosome system) may be involved in the disease pathogenesis. The autophagic system is a conserved intracellular system designed for the degradation of long-lived proteins and organelles in lysosomes (25, 26). Three types of autophagy have been described: macroautophagy, microautophagy, and chaperon-mediated autophagy. Whereas macroautophagy and microautophagy involve the “in bulk” degradation of regions of the cytosol (27, 28), chaperon-mediated autophagy is a more selective pathway, and only proteins with a lysosomal targeting sequence are degraded (29). Cumulative evidence has suggested that an age-dependent decrease in the autophagy-lysosome system may account for the accumulation of abnormal proteins during aging (30, 31).Macroautophagy is induced when an isolation membrane is formed surrounding cytosolic components, forming an autophagic vacuole, which will eventually fuse with lysosomes for protein/organelle degradation. Induction of the isolation membrane is negatively regulated by mTOR (mammalian target of rapamycin) (32). It has been shown that increasing autophagy activation by mTOR inhibitors has beneficial effects in neurodegeneration (3335).  相似文献   

5.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

6.
7.
8.
9.
Flap endonuclease 1 (FEN1) proteins, which are present in all kingdoms of life, catalyze the sequence-independent hydrolysis of the bifurcated nucleic acid intermediates formed during DNA replication and repair. How FEN1s have evolved to preferentially cleave flap structures is of great interest especially in light of studies wherein mice carrying a catalytically deficient FEN1 were predisposed to cancer. Structural studies of FEN1s from phage to human have shown that, although they share similar folds, the FEN1s of higher organisms contain a 3′-extrahelical nucleotide (3′-flap) binding pocket. When presented with 5′-flap substrates having a 3′-flap, archaeal and eukaryotic FEN1s display enhanced reaction rates and cleavage site specificity. To investigate the role of this interaction, a kinetic study of human FEN1 (hFEN1) employing well defined DNA substrates was conducted. The presence of a 3′-flap on substrates reduced Km and increased multiple- and single turnover rates of endonucleolytic hydrolysis at near physiological salt concentrations. Exonucleolytic and fork-gap-endonucleolytic reactions were also stimulated by the presence of a 3′-flap, and the absence of a 3′-flap from a 5′-flap substrate was more detrimental to hFEN1 activity than removal of the 5′-flap or introduction of a hairpin into the 5′-flap structure. hFEN1 reactions were predominantly rate-limited by product release regardless of the presence or absence of a 3′-flap. Furthermore, the identity of the stable enzyme product species was deduced from inhibition studies to be the 5′-phosphorylated product. Together the results indicate that the presence of a 3′-flap is the critical feature for efficient hFEN1 substrate recognition and catalysis.In eukaryotic DNA replication and repair, various bifurcated nucleic acid structure intermediates are formed and must be processed by the appropriate nuclease. Two examples of biological processes that create bifurcated DNA intermediates are Okazaki fragment maturation (1, 2) and long patch excision repair (3). In both models, a polymerase executes strand-displacement synthesis to create a double-stranded DNA (dsDNA)6 two-way junction from which a 5′-flap structure protrudes. The penultimate step of both pathways is the cleavage of this flap structure to create a nicked DNA that is then ligated. Because the bifurcated DNA structures that are formed in the aforementioned processes can theoretically occur anywhere in the genome, the nuclease associated with the cleavage of 5′-flap structures in eukaryotic cells, which is called flap endonuclease 1 (FEN1), must be capable of cleavage regardless of sequence. Therefore, FEN1 nucleases, which are found in all kingdoms of life (4), have evolved to recognize substrates based upon nucleic acid structure and strand polarity (5, 6).The Okazaki fragment maturation pathway of yeast has become a paradigm of eukaryotic lagging strand DNA synthesis. In the yeast model, bifurcated intermediates with large single-stranded DNA (ssDNA) 5′-flap structures are imprecisely cleaved by DNA2 in a replication protein A -dependent manner (7). Subsequent to the DNA2 cleavage, Rad27 (yeast homologue of FEN1) cleaves precisely to generate an intermediate suitable for ligation (2). The recent discovery that human DNA2 is predominantly located in mitochondria in various human cell lines (8, 9) suggests that hFEN1 is the paramount 5′-flap endonuclease in the nuclei of human cells. This observation potentially provides a plausible rationale for why deletion of RAD27 (yeast FEN1 homologue) is tolerated in Saccharomyces cerevisiae (10), whereas deletion of FEN1 in mammals is embryonically lethal (11). Recent models wherein mice carrying a mutation (E160D) in the FEN1 gene, which was shown in vitro to alter enzymatic properties (12), have demonstrated that FEN1 functional deficiency in mice (S129 and Black 6) increases the incidence of cancer, albeit different types presumably due to genetic background (13, 14). Thus, the function of mammalian FEN1 in vivo is vital to the prevention of genomic instability. In addition to its importance in the nucleus, hFEN1 has recently been detected in mitochondrial extracts (15, 16) and implicated in mitochondrial long patch base excision repair (15). Considering the pivotal roles of hFEN1 in DNA replication and repair, it is of interest to understand how hFEN1 and homologues achieve substrate and scissile phosphate selectivity in the absence of sequence information.Since its initial discovery as a nuclease that completes reconstituted Okazaki fragment maturation (17) and subsequent rediscovery as a 5′-flap-specific nuclease (DNaseIV) from bacteria (18), mouse (19), and HeLa cells (20), FEN1 proteins ranging from phage to human have been studied biochemically, computationally, and structurally (5, 6, 21). Biochemical characterizations of FEN1 proteins from various organisms have shown that this family of nucleases can perform phosphodiesterase activity on a wide variety of substrates; however, the efficiency of catalysis on various substrates differs among the species. For instance, phage FEN1s prefer pseudo-Y substrates (22, 23), whereas the archaeal and eukaryotic FEN1s prefer 5′-flap substrates (21, 24, 25), which have two dsDNA domains, one upstream and downstream of the site of cleavage, and a 5′-ssDNA protrusion (Fig. 1A). Primary sequence analysis indicates that FEN1 proteins share characteristic N-terminal (N) and Intermediate (I) “domains,” which harbor the highly conserved carboxylate residues that bind the requisite divalent metal ions (2628). Structural studies of FEN1 nucleases from phage to humans (22, 2936), have shown that the N and I domains comprise a single nuclease core domain consisting of a mixed, six- or seven-stranded β-sheet packed against an α-helical structure on both sides. The α-helices on either side of the β-sheet are “bridged” by a helical arch that spans the active site groove (supplemental Fig. S1). On one side of the β-sheet, the α-helical bundle (αb1) creates the floor of the active site and a DNA binding motif (helix-3-turn-helix) (32). Similarly, the opposite α-helical bundle (αb2) has also been observed to interact with DNA (35). Based on site-directed mutagenesis studies with T5 phage FEN1 (T5FEN1) (37) and hFEN1 (38, 39), and crystallographic studies of T4 phage FEN1 (T4FEN1) (22) and Archaeoglobus fulgidus FEN1 (aFEN1) (35) in complex with DNA, a general model for how FEN1 proteins recognize flap DNA has emerged. The helix-3-turn-helix motif is involved in downstream dsDNA binding, whereas the upstream dsDNA domain is bound by αb2. The helical arch is likely involved in 5′-flap binding (22).Open in a separate windowFIGURE 1.Secondary structure schematics of hFEN1 substrates. A, illustration of a general flap substrate created using a bimolecular approach whereby a template strand (T-strand), which partially folds into a hairpin, anneals with the duplex strand (d-strand). The T-strand hairpin creates the upstream dsDNA domain, whereas the d-strand base pairs with the T-strand to create the downstream dsDNA domain. The flap or any other structure is created by addition of nucleotides to the 5′-end of the d-strand. The interface between the upstream and downstream dsDNA domains may be viewed as a derivative of a two-way junction (74). Annealing of either the F(5), E, or G(15) d-strands with the T3F T-strand results in the formation of a (B) double flap substrate (Flap of 5-nt d-strand paired with a Template with a 3′-Flap, F(5)·T3F), C, exonuclease substrate with a 3′-extrahelical nucleotide (EXO d-strand paired with a Template with a 3′-Flap, E·T3F), and a D, fork-GEN substrate with a 3′-extrahelical nucleotide and a 15-nt ssDNA gap capped by a 23-nt hairpin structure (fork-Gap of 15-nt d-strand paired with a Template with a 3′-Flap, G(15)·T3F). E, annealing the F(5) d-strand with the T oligonucleotide creates a single flap (Flap of 5-nt d-strand paired with a Template, F(5)·T).Unlike phage FEN1s, studies of FEN1s from eubacterial (40), archaeal (21), and eukaryotic origins (41) have shown that the addition of a 3′-extrahelical nucleotide (3′-flap) to the upstream duplex of a 5′-flap substrate results in a rate enhancement and an increase in cleavage site specificity. Moreover, substrates possessing a 3′-flap, which mimic physiological “equilibrating flaps,” were cleaved exactly one nucleotide into the downstream duplex, thereby resulting in 5′-phosphorylated dsDNA product that was a suitable substrate for DNA ligase I (21, 41). As postulated by Kaiser et al. (21), the structure of an archaeal FEN1 in complex with dsDNA with a 3′-overhang showed that the protein contains a cleft adjacent to the upstream dsDNA binding site that binds the 3′-flap by means of van der Waals and hydrogen bonding interactions with the sugar moiety (35). Once the residues associated with 3′-flap binding were identified, sequence alignment analyses showed that the amino acid residues in the 3′-flap binding pocket are highly conserved from archaea to human. Furthermore, mutation of the conserved amino acid residues in the 3′-flap binding pocket of hFEN1 resulted in reduced affinity for and cleavage specificity on double flap substrates (42). Although the effects of the addition of a 3′-flap to substrates on hFEN1 catalysis are known qualitatively, a detailed understanding of the relationship between changes in catalytic parameters and rate enhancement by the presence of a 3′-flap is unknown. Here, we describe a detailed kinetic analysis of hFEN1 using four well characterized DNA substrates and show that the presence of a 3′-flap on a substrate not only contributes to substrate binding (42), but also increases multiple and single turnover rates of reaction in the presence of near physiological monovalent salt concentrations. We also demonstrate that, like T5FEN1, hFEN1 is rate-limited by product release, and thus multiple turnover rates at saturating concentrations of substrate are predominantly a reflection of product release and not catalysis as was previously concluded (39). Furthermore, this study provides insight into the mechanism of hFEN1 substrate recognition.  相似文献   

10.
Plasma proteome analysis requires sufficient power to compare numerous samples and detect changes in protein modification, because the protein content of human samples varies significantly among individuals, and many plasma proteins undergo changes in the bloodstream. A label-free proteomics platform developed in our laboratory, termed “Two-Dimensional Image Converted Analysis of Liquid chromatography and mass spectrometry (2DICAL),” is capable of these tasks. Here, we describe successful detection of novel prolyl hydroxylation of α-fibrinogen using 2DICAL, based on comparison of plasma samples of 38 pancreatic cancer patients and 39 healthy subjects. Using a newly generated monoclonal antibody 11A5, we confirmed the increase in prolyl-hydroxylated α-fibrinogen plasma levels and identified prolyl 4-hydroxylase A1 as a key enzyme for the modification. Competitive enzyme-linked immunosorbent assay of 685 blood samples revealed dynamic changes in prolyl-hydroxylated α-fibrinogen plasma level depending on clinical status. Prolyl-hydroxylated α-fibrinogen is presumably controlled by multiple biological mechanisms, which remain to be clarified in future studies.For comprehensive analysis of plasma proteins, it is necessary to compare a sufficient number of blood samples to avoid simple interindividual heterogeneity, because the protein content of human samples varies significantly among individuals. Also, the provision of sufficient power is needed to detect protein modification because many plasma proteins undergo changes in the bloodstream (1). Even though the proteomic technologies have advanced (2, 3), there remains room for improvement. Different isotope labeling and identification-based methods have been developed for quantitative proteomics technologies (46), but the number of samples that can be compared by the current isotope-labeling methods is limited, and identification-based proteomics is unable to capture information regarding unknown modifications.A label-free proteomics platform developed in our laboratory, termed “Two-Dimensional Image Converted Analysis of Liquid chromatography and mass spectrometry (2DICAL)2 (7), simply compares the liquid chromatography and mass spectrometry (LC-MS) data and detects a protein modification by finding changes in the mass to charge ratio (m/z) and retention time (RT). Enhanced methods for accurate MS peak alignment across multiple LC runs have enabled the successful implementation of clinical studies requiring comparison of a large number of samples (8, 9). Using 2DICAL to analyze plasma samples of pancreatic cancer patients and healthy controls, novel prolyl hydroxylation of α-fibrinogen was successfully discovered.Fibrinogen and its modification has been investigated because of its clinical importance (10, 11). On the other hand, prolyl hydroxylation has attracted attention after the discovery of the hypoxia-inducible factor 1α (HIF1α) prolyl-hydroxylase and its role in switching of HIF1α functions (12). Prolyl hydroxylation in other proteins has been energetically sought, but only a few such proteins have been identified (13). Only one study has reported prolyl hydroxylation of fibrinogen at the β chain (14).Here, we report the detection of prolyl 4-hydroxylated α-fibrinogen by plasma proteome analysis, a protein modification that dynamically changes in plasma depending on the clinical status and is a candidate plasma biomarker.  相似文献   

11.
12.
We outline several principles that we believe define the gating of two bacterial mechanosensitive channels, MscL and MscS. Serving as turgor regulators in bacteria and other walled cells, these molecules are tangible models for studying conformational transitions in membrane proteins driven directly by membrane tension. MscL, a compact pentamer, reversibly opens a gigantic 30-Å pore at near-lytic tensions. MscS, a heptameric complex, exhibits transient activation of a smaller pore at moderate tensions, thereby entering a tension-insensitive inactivated state. By comparing the structures and predicted transitions in these channels, we concluded that opening is commonly achieved through tilting and outward motion of the pore-lining helices, which is kinetically limited by hydration of the pore. The intricate adaptive behavior in MscS appears to depend on specific interhelical associations and the flexibility of the pore-lining helices. We discuss physical factors that may direct the transitions and stabilize main functional states in these channels.Osmotic forces are strong, which necessitated development of osmoregulation along with the first semipermeable membrane delineating the early cell. A simple estimation shows that a 1-μm cell behaving as an ideal osmometer would sustain a downshock no stronger than 20 mm, after which membrane tension would exceed the lytic limit of 10–12 dynes/cm. Thus, a cell without a reinforcing envelope or protective valves is very vulnerable. Free-living and enteric microorganisms cycling through the soil and experiencing drastic environmental changes developed robust mechanisms to maintain volume and integrity (1). The mechanosensitive channels MscS and MscL (mechanosensitive channels of small and large conductance, respectively) have been identified as primary osmolyte release valves limiting the turgor pressure under acute osmotic shock (24).Without mscS and mscL genes, Escherichia coli survives a 300 mosm osmotic downshock (2), its resistance attributed to the peptidoglycan layer partially restraining swelling. However, expression of either MscS or MscL allows cells to withstand a 700–800 mosm downshock through release of small osmolytes (2). Purification and reconstitution proved that MscL and MscS respond directly to tension in the lipid bilayer (57). Both channels reside in the inner (cytoplasmic) membrane (8), with MscL localized at the cell poles, bearing high curvature (9).As primary components of the turgor regulation system, E. coli MscS and MscL became convenient models for studies of tension-driven conformational transitions in membrane proteins (10). The crystal structures of closed-state Mycobacterium tuberculosis MscL (11) and E. coli MscS in two distinct conformations (12, 13) provided invaluable initial points to explore their gating mechanisms, in which computational methods play increasingly important roles.  相似文献   

13.
The binding of the adaptor protein APPL1 to adiponectin receptors is necessary for adiponectin-induced AMP-activated protein kinase (AMPK) activation in muscle, yet the underlying molecular mechanism remains unknown. Here we show that in muscle cells adiponectin and metformin induce AMPK activation by promoting APPL1-dependent LKB1 cytosolic translocation. APPL1 mediates adiponectin signaling by directly interacting with adiponectin receptors and enhances LKB1 cytosolic localization by anchoring this kinase in the cytosol. Adiponectin also activates another AMPK upstream kinase Ca2+/calmodulin-dependent protein kinase kinase by activating phospholipase C and subsequently inducing Ca2+ release from the endoplasmic reticulum, which plays a minor role in AMPK activation. Our results show that in muscle cells adiponectin is able to activate AMPK via two distinct mechanisms as follows: a major pathway (the APPL1/LKB1-dependent pathway) that promotes the cytosolic localization of LKB1 and a minor pathway (the phospholipase C/Ca2+/Ca2+/calmodulin-dependent protein kinase kinase-dependent pathway) that stimulates Ca2+ release from intracellular stores.Adiponectin, an adipokine abundantly expressed in adipose tissue, exhibits anti-diabetic, anti-inflammatory, and anti-atherogenic properties and hence is a potential therapeutic target for various metabolic diseases (13). The beneficial effects of adiponectin are mediated through the direct interaction of adiponectin with its cell surface receptors, AdipoR1 and AdipoR2 (4, 5). Adiponectin increases fatty acid oxidation and glucose uptake in muscle cells by activating AMP-activated protein kinase (AMPK)3 (4, 6), which depends on the interaction of AdipoR1 with the adaptor protein APPL1 (Adaptor protein containing Pleckstrin homology domain, Phosphotyrosine binding domain, and Leucine zipper motif) (5). However, the underlying mechanisms by which APPL1 mediates adiponectin signaling to AMPK activation and other downstream targets remain unclear.AMPK is a serine/threonine protein kinase that acts as a master sensor of cellular energy balance in mammalian cells by regulating glucose and lipid metabolism (7, 8). AMPK is composed of a catalytic α subunit and two noncatalytic regulatory subunits, β and γ. The NH2-terminal catalytic domain of the AMPKα subunit is highly conserved and contains the activating phosphorylation site (Thr172) (9). Two AMPK variants, α1 and α2, exist in mammalian cells that show different localization patterns. AMPKα1 subunit is localized in non-nuclear fractions, whereas the AMPKα2 subunit is found in both nucleus and non-nuclear fractions (10). Biochemical regulation of AMPK activation occurs through various mechanisms. An increase in AMP level stimulates the binding of AMP to the γ subunit, which induces a conformational change in the AMPK heterotrimer and results in AMPK activation (11). Studies have shown that the increase in AMPK activity is not solely via AMP-dependent conformational change, rather via phosphorylation by upstream kinases, LKB1 and CaMKK. Dephosphorylation by protein phosphatases is also important in regulating the activity of AMPK (12).LKB1 has been considered as a constitutively active serine/threonine protein kinase that is ubiquitously expressed in all tissues (13, 14). Under conditions of high cellular energy stress, LKB1 acts as the primary AMPK kinase through an AMP-dependent mechanism (1517). Under normal physiological conditions, LKB1 is predominantly localized in the nucleus. LKB1 is translocated to the cytosol, either by forming a heterotrimeric complex with Ste20-related adaptor protein (STRADα/β) and mouse protein 25 (MO25α/β) or by associating with an LKB1-interacting protein (LIP1), to exert its biological function (1822). Although LKB1 has been shown to mediate contraction- and adiponectin-induced activation of AMPK in muscle cells, the underlying molecular mechanisms remain elusive (15, 23).CaMKK is another upstream kinase of AMPK, which shows considerable sequence and structural homology with LKB1 (2426). The two isoforms of CaMKK, CaMKKα and CaMKKβ, encoded by two distinct genes, share ∼70% homology at the amino acid sequence level and exhibit a wide expression in rodent tissues, including skeletal muscle (2734). Unlike LKB1, AMPK phosphorylation mediated by CaMKKs is independent of AMP and is dependent only on Ca2+/calmodulin (35). Hence, it is possible that an LKB1-independent activation of AMPK by CaMKK exists in muscle cells. However, whether and how adiponectin stimulates this pathway in muscle cells are not known.In this study, we demonstrate that in muscle cells adiponectin induces an APPL1-dependent LKB1 translocation from the nucleus to the cytosol, leading to increased AMPK activation. Adiponectin also activates CaMKK by stimulating intracellular Ca2+ release via the PLC-dependent mechanism, which plays a minor role in activation of AMPK. Taken together, our results demonstrate that enhanced cytosolic localization of LKB1 and Ca2+-induced activation of CaMKK are the mechanisms underlying adiponectin-stimulated AMPK activation in muscle cells.  相似文献   

14.
Group I metabotropic glutamate receptors (mGluRs) are coupled via phospholipase Cβ to the hydrolysis of phosphoinositides and function to modulate neuronal excitability and synaptic transmission at glutamatergic synapses. The desensitization of Group I mGluR signaling is thought to be mediated primarily via second messenger-dependent protein kinases and G protein-coupled receptor kinases. We show here that both mGluR1 and mGluR5 interact with the calcineurin inhibitor protein (CAIN). CAIN is co-immunoprecipitated in a complex with Group I mGluRs from both HEK 293 cells and mouse cortical brain lysates. Purified CAIN and its C-terminal domain specifically interact with glutathione S-transferase fusion proteins corresponding to the second intracellular loop and the distal C-terminal tail domains of mGluR1. The interaction of CAIN with mGluR1 could also be blocked using a Tat-tagged peptide corresponding to the mGluR1 second intracellular loop domain. Overexpression of full-length CAIN attenuates the agonist-stimulated endocytosis of both mGluR1a and mGluR5a in HEK 293 cells, but expression of the CAIN C-terminal domain does not alter mGluR5a internalization. In contrast, overexpression of either full-length CAIN or the CAIN C-terminal domain impairs agonist-stimulated inositol phosphate formation in HEK 293 cells expressing mGluR1a. This CAIN-mediated antagonism of mGluR1a signaling appears to involve the disruption of receptor-Gαq/11 complexes. Taken together, these observations suggest that the association of CAIN with intracellular domains involved in mGluR/G protein coupling provides an additional mechanism by which Group I mGluR endocytosis and signaling are regulated.Metabotropic glutamate receptors (mGluRs)2 play an essential role in regulating neuronal plasticity, development, and neurotoxicity and belong to the G protein-coupled receptor superfamily of integral membrane proteins (14). The mGluR family can be subclassified into three groups based on sequence homology, G protein specificity, and pharmacology. Group I mGluRs (mGluR1 and mGluR5) couple via the heterotrimeric Gαq/11 proteins to the activation of phospholipase Cβ, resulting in the formation of inositol 1,4,5-triphosphate and diacylglycerol, the release of Ca2+ from intracellular stores, and the activation of protein kinase C (PKC) (46).The regulation of mGluR signal transduction involves numerous proteins that function to regulate signaling at both the level of the heterotrimeric G protein and the receptor (68). At the level of the receptor, Group I mGluR activity is regulated by a process termed desensitization, which protects against both acute and chronic receptor overstimulation (9, 10). The attenuation of Group I mGluR signaling can be mediated by both phosphorylation-dependent and phosphorylation-independent processes (11). The phosphorylation-independent attenuation of Group I mGluR signaling is mediated by GRK2 (G protein-coupled receptor kinase 2), which is composed of three functional domains: an N-terminal RGS (regulator of G protein signaling) homology domain, a central catalytic domain, and a C-terminal Gβγ-binding pleckstrin homology domain (12). GRK2-mediated desensitization of Group I mGluRs does not require catalytic activity but rather requires the interaction of the GRK2 RGS homology domain with both the second intracellular loop domain of mGluR1 and the α-subunit of Gαq/11, thereby attenuating heterotrimeric G protein coupling (1315). Phosphorylation-independent desensitization of mGluR1 signaling is also mediated by optineurin, an effect that is enhanced by the expression of mutant huntingtin (16). Phosphorylation-dependent desensitization of Group I mGluR responsiveness involves the phosphorylation of PKC consensus sequence localized within the intracellular loop and C-terminal tail domains of mGluR1 and mGluR5 by PKC (17, 18). It is proposed that calcineurin and mGluR5 may exist in a signaling complex in the brain and that calcineurin may function to modulate mGluR5 signaling by directly dephosphorylating the receptor at a PKC consensus site that contributes to mGluR5 desensitization (19). Calcineurin is also linked to the regulation of endocytosis via its interaction with dynamin-1 (20).On the basis of the observation that calcineurin may form a complex with Group I mGluRs, we hypothesized that CAIN (calcineurin inhibitor protein) might also interact with Group I mGluRs and modulate their endocytosis and signaling. CAIN, also known as Cabin1 (calcineurin-binding protein), was first identified as a protein that binds to calcineurin and was shown to inhibit calcineurin catalytic activity (2123). Previous studies also demonstrated that CAIN may interact with amphiphysin-1, dynamin-1, and α-adaptin and led to the suggestion that CAIN functions as a component of synaptic endocytic complexes (24). Consistent with this hypothesis, the overexpression of CAIN in human embryonic kidney (HEK 293) cells resulted in attenuated transferrin receptor endocytosis.We show here that CAIN interacts with the second intracellular loop and C-terminal tail domains of Group I mGluRs, inhibits Group I mGluR internalization, and attenuates mGluR1a signaling by disrupting receptor-Gαq/11 complexes. Taken together, these results describe an additional mechanism by which Group I mGluR activity may be regulated.  相似文献   

15.
16.
Cells exposed to environmental stress rapidly activate the MAPK cascade (MKKK/MKK/MAPK). The transient nature of stress signaling is a consequence of negative feedback signals that lead to kinase dephosphorylation, degradation, and sequestration, which have not been fully elucidated for MKK family members. Here, we investigated the signals that negatively regulate MKK4/SEK1, an upstream activator of the MAPKs JNK and p38/HOG1. Following exposure of cells to sorbitol, MKK4 underwent ubiquitination and degradation in a proteasome-dependent manner. MKK4 ubiquitination required JNK kinase activity. The JNK substrate Itch (a HECT domain-containing Nedd4-like ubiquitin protein ligase) bound to MKK4, ubiquitinated lysines 140 and 143, and promoted MKK4 degradation. Other E3 ligases within the MAPK modular complex did not ubiquitinate MKK4. These data suggest that MKK4 is negatively regulated through a feedback loop involving the E3 ubiquitin ligase Itch, which has a fundamental role in the mechanism that controls MKK4 protein levels.Cellular responses to environmental stress are regulated by an intracellular phosphorelay system involving at least four groups of MAPKs,2 including ERK1/2, JNK1/2/3, p38α/β/γ/δ, and ERK5, which are substrates for specific MAP2K proteins (MKKs); ERK1 and ERK2 are substrates for MKK1/2, p38 for MKK3/6, JNK for MKK4/7, and ERK5 for MKK5 (1, 2). Each MKK is regulated by multiple MAP3K proteins (MKKKs), of which there are more than a dozen mammalian family members, increasing the complexity and diversity of MAPK signaling (35). The specificity of these protein interactions is preserved by scaffolding proteins that organize into a single module composed of the three components of a MAPK cascade and its upstream activators (1, 6).Cells exposed in vitro to stress typically exhibit rapid activation and decay of MAPK activity (2). The degradation of the MAPK signal is a consequence of negative feedback loops that regulate kinase activity, abundance, and localization through changes in kinase phosphorylation and ubiquitination (3, 6, 7). For example, kinase phosphorylation regulates interactions with E3 ligases, which transfer polyubiquitin chains onto lysine residues, by regulating the subcellular location of the kinase or by creating phosphodegrons, which are recognition signals for specific E3 ligases (810). The transfer of ubiquitin occurs passively in the case of RING (really interesting new gene) finger-containing E3 ligases, which function as adaptor molecules between the E2 ubiquitin-conjugating enzyme and substrate, or actively in the case of HECT (homologous to the E6-associated protein C terminus) domain-containing E3 ligases, which serve as a catalytic intermediate in the transfer process (11, 12). Through their interactions with E3 ligases, certain MKKKs (MEKK1 and MEKK2) (13, 14) and MAPKs (ERK2 and ERK7) (15, 16) undergo ubiquitination, which marks these proteins for degradation by the 26 S proteasome, thereby attenuating MAPK signaling. In contrast, less progress has been made in understanding how mammalian MKK family members are negatively regulated. Given that MKK homologs in yeast (17) and Dictyostelium (18) are ubiquitinated following prolonged stimulation, we postulated that, like MKKKs and MAPKs, mammalian MKKs are regulated through protein ubiquitination, a reasonable postulate given the modular organization and coordinate regulation of kinases within the MAPK cascade.In this study, we found that MKKs undergo ubiquitination and proteasomal degradation in response to environmental stress. MKK4 associates with the ubiquitin ligase Itch (which belongs to the HECT domain-containing Nedd4-like E3 family), is an important regulator of murine epithelial and hematopoietic cell function (19), and is absent in 18H (Itchy) mice, which exhibit profound immune defects (20). Itch binds to and ubiquitinates MKK4 and mediates MKK4 protein degradation. Notably, stress-induced MKK4 degradation is dependent upon activation of the MKK4 substrate JNK, which phosphorylates and activates Itch. We conclude that MKK4 protein stability is regulated through a negative feedback loop involving Itch.  相似文献   

17.
Proteolytic processing modifies the pleiotropic functions of many large, complex, and modular proteins and can generate cleavage products with new biological activity. The identification of exact proteolytic cleavage sites in the extracellular matrix laminins, fibronectin, and other extracellular matrix proteins is not only important for understanding protein turnover but is needed for the identification of new bioactive cleavage products. Several such products have recently been recognized that are suggested to play important cellular regulatory roles in processes, including angiogenesis. However, identifying multiple cleavage sites in extracellular matrix proteins and other large proteins is challenging as N-terminal Edman sequencing of multiple and often closely spaced cleavage fragments on SDS-PAGE gels is difficult, thus limiting throughput and coverage. We developed a new liquid chromatography-mass spectrometry approach we call amino-terminal oriented mass spectrometry of substrates (ATOMS) for the N-terminal identification of protein cleavage fragments in solution. ATOMS utilizes efficient and low cost dimethylation isotopic labeling of original N-terminal and proteolytically generated N termini of protein cleavage fragments followed by quantitative tandem mass spectrometry analysis. Being a peptide-centric approach, ATOMS is not dependent on the SDS-PAGE resolution limits for protein fragments of similar mass. We demonstrate that ATOMS reliably identifies multiple proteolytic sites per reaction in complex proteins. Fifty-five neutrophil elastase cleavage sites were identified in laminin-1 and fibronectin-1 with 34 more identified by matrix metalloproteinase cleavage. Hence, our degradomics approach offers a complimentary alternative to Edman sequencing with broad applicability in identifying N termini such as cleavage sites in complex high molecular weight extracellular matrix proteins after in vitro cleavage assays. ATOMS can therefore be useful in identifying new cleavage products of extracellular matrix proteins cleaved by proteases in pathology for bioactivity screening.Recently, considerable efforts have been deployed to develop high throughput proteomic screens to identify protease substrates in complex biological samples (18). Validation of substrates identified by these approaches or identification of cleavage sites by in vitro incubation of candidate substrates with the protease of interest is generally performed by SDS-PAGE analysis and Edman degradation and sequencing. However, the complexity of large modular proteins renders Edman sequencing of proteolytic fragments difficult to apply because each of the numerous proteolytic fragments should be analyzed separately, and high coverage of cleavage sites is rarely attained (9). Cleavage site identification after protein degradation is also very difficult for small peptide products less than 4 kDa. Consequently, the precise cleavage sites in complex extracellular matrix proteins such as laminin and fibronectin by important tissue and inflammatory cell proteases such as the matrix metalloproteinases (MMPs)1 and neutrophil elastase are mostly unknown.These limitations of Edman sequencing are problematic in the study of tissue remodeling and proteolysis in pathology. Neutrophil elastase and several MMPs such as MMP2, MMP8, and MMP9 play key roles in inflammation (10, 11), tissue healing (12, 13), and carcinogenesis (14, 15) and are well known for degrading extracellular matrix proteins (16). More recently, signaling functions for MMPs are increasingly recognized as one of their most important roles by the precise processing of cytokines or their binding proteins (17). In addition, several important examples are now known of cryptic binding sites being exposed after precise protein cleavage or new proteins termed neoproteins (18) being released upon limited cleavage of extracellular matrix proteins and having completely different functions compared with their parent molecule, including several with importance in angiogenesis (1925). Many such sites or neoproteins are generated by inflammatory proteases or proteases of the coagulation and fibrinolysis systems (24, 25), and this is a burgeoning field of discovery that is often hampered by difficulties in their N-terminal sequencing.In light of this limitation, we developed, validated, and used a new method for targeted and simultaneous N-terminal sequencing of one or a small number of protein N termini or cleavage products we call amino-terminal oriented mass spectrometry of substrates (ATOMS). We applied ATOMS for the analysis of cleavage sites generated in laminin-1 and fibronectin-1 by neutrophil elastase and neutrophil and tissue MMPs. Laminin-1 (LM-111), a trimeric glycoprotein composed of the α1, β1, and γ1 chains, is ubiquitously expressed in epithelium and endothelium. Proteolytic processing of laminins greatly affects cellular behavior and is also implicated in cancer cell migration (20, 2629). Another important extracellular matrix protein is plasma fibronectin (also known as fibronectin isoform 1) and its cellular isoforms, which are homodimers linked by a disulfide bridge at the C terminus (30) that are important for cell adhesion and intracellular signaling (3134). Fibronectin is susceptible to proteolysis (35, 36), which affects its biological functions (3739). However, the cleavage sites within these two molecules by inflammatory MMPs and neutrophil elastase are largely unknown. Here we identified a total of 55 neutrophil elastase cleavage sites in LM-111 and fibronectin-1 and 34 MMP cleavage sites, demonstrating the capacity of ATOMS to identify multiple N-terminal sequences in solution. ATOMS also outperformed N-terminal Edman sequencing with 50% more cleavage sites identified by ATOMS, representing a significant advance in N-terminal sequencing technology. The utility of the method is broadly applicable for the analysis of multiple cleavages in other very large molecules and so offers great potential to accurately identify and rapidly sequence multiple cryptic bioactive protein fragments liberated following proteolytic processing.  相似文献   

18.
Caspase-3 and -7 are considered functionally redundant proteases with similar proteolytic specificities. We performed a proteome-wide screen on a mouse macrophage lysate using the N-terminal combined fractional diagonal chromatography technology and identified 46 shared, three caspase-3-specific, and six caspase-7-specific cleavage sites. Further analysis of these cleavage sites and substitution mutation experiments revealed that for certain cleavage sites a lysine at the P5 position contributes to the discrimination between caspase-7 and -3 specificity. One of the caspase-7-specific substrates, the 40 S ribosomal protein S18, was studied in detail. The RPS18-derived P6–P5′ undecapeptide retained complete specificity for caspase-7. The corresponding P6–P1 hexapeptide still displayed caspase-7 preference but lost strict specificity, suggesting that P′ residues are additionally required for caspase-7-specific cleavage. Analysis of truncated peptide mutants revealed that in the case of RPS18 the P4–P1 residues constitute the core cleavage site but that P6, P5, P2′, and P3′ residues critically contribute to caspase-7 specificity. Interestingly, specific cleavage by caspase-7 relies on excluding recognition by caspase-3 and not on increasing binding for caspase-7.Caspases, a family of evolutionarily conserved proteases, mediate apoptosis, inflammation, proliferation, and differentiation by cleaving many cellular substrates (13). The apoptotic initiator caspases (caspase-8, -9, and -10) are activated in large signaling platforms and propagate the death signal by cleavage-induced activation of executioner caspase-3 and -7 (4, 5). Most of the cleavage events occurring during apoptosis have been attributed to the proteolytic activity of these two executioner caspases, which can act on several hundreds of proteins (2, 3, 6, 7). The substrate degradomes of the two main executioner caspases have not been determined but their identification is important to gaining greater insight in their cleavage specificity and biological functions.The specificity of caspases was rigorously profiled by using combinatorial tetrapeptide libraries (8), proteome-derived peptide libraries (9), and sets of individual peptide substrates (10, 11). The results of these studies indicate that specificity motifs for caspase-3 and -7 are nearly indistinguishable with the canonical peptide substrate, DEVD, used to monitor the enzymatic activity of both caspase-3 and -7 in biological samples. This overlap in cleavage specificity is manifested in their generation of similar cleavage fragments from a variety of apoptosis-related substrates such as inhibitor of caspase-activated DNase, keratin 18, PARP,1 protein-disulfide isomerase, and Rho kinase I (for reviews, see Refs. 2, 3, and 7). This propagated the view that these two caspases have completely redundant functions during apoptosis. Surprisingly, mice deficient in one of these caspases (as well as mice deficient in both) have distinct phenotypes. Depending on the genetic background of the mice, caspase-3-deficient mice either die before birth (129/SvJ) or develop almost normally (C57BL/6J) (1214). This suggests that dynamics in the genetic background, such as increased caspase-7 expression, compensate for the functional loss of caspase-3 (15). In the C57BL/6J background, caspase-7 single deficient mice are also viable, whereas caspase-3 and -7 double deficient mice die as embryos, further suggesting redundancy (1214). However, because caspase-3 and -7 probably arose from gene duplication between the Cephalochordata-Vertebrata diversion (16), they might have acquired different substrate specificities during evolution. Caspase-3 and -7 do exhibit different activities on a few arbitrarily identified natural substrates, including BID, X-linked inhibitor of apoptosis protein, gelsolin, caspase-6, ataxin-7, and co-chaperone p23 (1720). In addition, caspase-3 generally cleaves more substrates during apoptosis than caspase-7 and therefore appears to be the major executioner caspase. Moreover, a recent report describing caspase-1-dependent activation of caspase-7, but not of caspase-3, in macrophages in response to microbial stimuli supports the idea of a non-redundant function for caspase-7 downstream of caspase-1 (21).Commercially available “caspase-specific” tetrapeptide substrates are widely used for specific caspase detection, but they display substantial promiscuity and cannot be used to monitor individual caspases in cells (22, 23). Detecting proteolysis by measuring the release of C-terminal fluorophores, such as 7-amino-4-methylcoumarin (amc), restricts the specificity of these peptide substrates to non-prime cleavage site residues, which may have hampered the identification of specific cleavage events. To address this limitation, a recently developed proteomics technique, called proteomic identification of protease cleavage sites, was used to map both non-prime and prime preferences for caspase-3 and -7 on a tryptic peptide library (9). However, no clear distinction in peptide recognition motifs between caspase-3 and -7 could be observed (9). Because not all classical caspase cleavage sites are processed (7), structural or post-translational higher order constraints are likely involved in steering the cleavage site selectivity. Peptide-based approaches generally overlook such aspects.We made use of the COFRADIC N-terminal peptide sorting methodology (2426) to profile proteolytic events of caspase-3 and -7 in a macrophage proteome labeled by triple stable isotope labeling by amino acids in cell culture (SILAC), which allowed direct comparison of peak intensities in peptide MS spectra and consequent quantification of N termini that are equally, preferably, or exclusively generated by the action of caspase-3 or -7 (26, 27). We identified 55 cleavage sites in 48 protein substrates, encompassing mutual, preferred, and unique caspase-3 and -7 cleavage sites.  相似文献   

19.
A reversibly glycosylated polypeptide from pea (Pisum sativum) is thought to have a role in the biosynthesis of hemicellulosic polysaccharides. We have investigated this hypothesis by isolating a cDNA clone encoding a homolog of Arabidopsis thaliana, Reversibly Glycosylated Polypeptide-1 (AtRGP1), and preparing antibodies against the protein encoded by this gene. Polyclonal antibodies detect homologs in both dicot and monocot species. The patterns of expression and intracellular localization of the protein were examined. AtRGP1 protein and RNA concentration are highest in roots and suspension-cultured cells. Localization of the protein shows it to be mostly soluble but also peripherally associated with membranes. We confirmed that AtRGP1 produced in Escherichia coli could be reversibly glycosylated using UDP-glucose and UDP-galactose as substrates. Possible sites for UDP-sugar binding and glycosylation are discussed. Our results are consistent with a role for this reversibly glycosylated polypeptide in cell wall biosynthesis, although its precise role is still unknown.The primary cell wall of dicot plants is laid down by young cells prior to the cessation of elongation and secondary wall deposition. Making up to 90% of the cell''s dry weight, the extracellular matrix is important for many processes, including morphogenesis, growth, disease resistance, recognition, signaling, digestibility, nutrition, and decay. The composition of the cell wall has been extensively described (Bacic et al., 1988; Levy and Staehelin, 1992; Zablackis et al., 1995), and yet many questions remain unanswered regarding the synthesis and interaction of these components to provide cells with a functional wall (Carpita and Gibeaut, 1993; Carpita et al., 1996).Heteropolysaccharide biosynthesis can be divided into four steps: (a) chain or backbone initiation, (b) elongation, (c) side-chain addition, and (d) termination and extracellular deposition (Waldron and Brett, 1985). The similarity between various polysaccharide backbones leads to the prediction that the synthesizing machinery would be conserved between them. For example, the backbone of xyloglucan polymers, β-1,4 glucan, can be synthesized independently of or concurrently with side-chain addition (Campbell et al., 1988; White et al., 1993), and this polymer and the chains that make up cellulose are identical. The later addition of side chains to xyloglucan are catalyzed by specific transferases (Kleene and Berger, 1993) such as xylosyltransferase (Campbell et al., 1988), galactosyltransferase, and fucosyltransferase (Faïk et al., 1997), all of which are localized to the Golgi compartment (Brummell et al., 1990; Driouich et al., 1993; Staehelin and Moore, 1995).The enzymes involved in wall biosynthesis have been recalcitrant to isolation (Carpita et al., 1996; Albersheim et al., 1997). Only recently has the first gene encoding putative cellulose biosynthetic enzymes, celA, been isolated from cotton (Gossypium hirsutum) and rice (Oryza sativa; Pear et al., 1996).During studies of polysaccharide synthesis in pea (Pisum sativum) Golgi membranes, Dhugga et al. (1991) identified a 41-kD protein doublet that they suggested was involved in polysaccharide synthesis. The authors showed that this protein could be glycosylated by radiolabeled UDP-Glc but that this labeling could be reversibly competed with by unlabeled UDP-Glc, UDP-Xyl, and UDP-Gal, the sugars that make up xyloglucan (Hayashi, 1989). The 41-kD protein was named PsRGP1 (P. sativum Reversibly Glycosylated Polypeptide-1; Dhugga et al., 1997). Furthermore, the conditions that stimulate or inhibit Golgi-localized β-glucan synthase activity are the same conditions that stimulate or inhibit the glycosylation of PsRGP1 (Dhugga et al., 1991). To address the role of this protein in polysaccharide synthesis, the authors purified the polypeptides and obtained the sequences from tryptic peptides (Dhugga and Ray, 1994). Antibodies raised against PsRGP1 showed that it is soluble and localized to the plasma membrane (Dhugga et al., 1991) and Golgi compartment (Dhugga et al., 1997). In addition to its Golgi localization, the steady-state glycosylation of PsRGP1 is approximately 10:7:3 (UDP-Glc:-Xyl:-Gal), which is similar to the typical sugar composition of xyloglucan (1.0:0.75:0.25; Dhugga et al., 1997).We were interested in studying various aspects of cell wall metabolism, including the synthesis of polysaccharides and their delivery to the cell wall. Studies in pea have shown that a 41-kD protein may be involved in cell wall polysaccharide synthesis, possibly that of xyloglucan (Dhugga et al., 1997). Here we report the characterization of AtRGP1 (Arabidopsis thaliana Reversibly Glycosylated Polypeptide-1), a soluble protein that can also be found weakly associated with membrane fractions, most likely the Golgi fraction. The reversible nature of the glycosylation of this Arabidopsis homolog by the substrates used to make polysaccharides (nucleotide sugars) suggests a possible role for AtRGP1 in polysaccharide biosynthesis.  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