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 31 毫秒
1.
2.
We have identified, purified, and characterized three subcomplexes of PA700, the 19 S regulatory complex of the 26 S proteasome. These subcomplexes (denoted PS-1, PS-2, and PS-3) collectively account for all subunits present in purified PA700 but contain no overlapping components or significant levels of non-PA700 proteins. Each subcomplex contained two of the six AAA subunits (Rpt1–6) that form the binding interface of PA700 with the 20 S proteasome, the protease component of the 26 S proteasome. Unlike intact PA700, no individual PA700 subcomplex displayed ATPase activity or proteasome activating activity. However, both activities were manifested by ATP-dependent in vitro reconstitution of PA700 from the subcomplexes. We exploited functional reconstitution to define and distinguish roles of different PA700 subunits in PA700 function by selective alteration of subunits within individual subcomplexes prior to reconstitution. Carboxypeptidase treatment of either PS-2 or PS-3, subcomplexes containing specific Rpt subunits previously shown to have important roles in 26 S proteasome assembly and activation, inhibited these processes but did not affect PA700 reconstitution or ATPase activity. Thus, the intact C termini of both subunits are required for 26 S proteasome assembly and activation but not for PA700 reconstitution. Surprisingly, carboxypeptidase treatment of PS-1 also inhibited 26 S proteasome assembly and activation upon reconstitution with untreated PS-2 and PS-3. These results suggest a previously unidentified role for other PA700 subunits in 26 S proteasome assembly and activation. Our results reveal relative structural and functional relationships among the AAA subunits of PA700 and new insights about mechanisms of 26 S proteasome assembly and activation.The 26 S proteasome is a 2,500,000-Da protease complex that degrades polyubiquitylated proteins by an ATP-dependent mechanism (1, 2). The biochemical processes required for this function are divided between two subcomplexes that compose the holoenzyme (3, 4). The first, called 20 S proteasome or core particle, is a 700,000-Da complex that catalyzes peptide bond hydrolysis (5). The second, called PA700 or 19 S regulatory particle, is a 700,000-Da complex that mediates multiple aspects of proteasome function related to initial binding and subsequent delivery of substrates to the catalytic sites of the 20 S proteasome (6). The 20 S proteasome is composed of 28 subunits representing the products of 14 genes arranged in four axially stacked heteroheptameric rings (7, 8). Each of the two center β rings contains three different protease subunits that utilize N-terminal threonine residues as catalytic nucleophiles (5, 8, 9). These residues line an interior lumen formed by the stacked rings and thus are sequestered from interaction with substrates by a shell of 20 S proteasome subunits.PA700 is composed of 20 different subunits. Six of these subunits, termed Rpt1–6, are AAA2 (ATPases Associated with various cellular Activities) family members that confer ATPase activity to the complex and mediate energy-dependent proteolysis by the 26 S proteasome (2, 10). 26 S proteasome assembly from PA700 and 20 S proteasome requires ATP binding to Rpt subunits (1115). Binding of PA700 to the 20 S proteasome occurs at an axial interface between a heterohexameric ring of the PA700 Rpt subunits and the heteroheptameric outer ring of α-type 20 S proteasome subunits (16). Substrates enter the proteasome through a pore in the center of the α subunit ring that is reversibly gated by conformationally variable N-terminal residues of certain α subunits in response to PA700 binding (12, 1719). Although the degradation of polyubiquitylated proteins requires additional ATP hydrolysis-dependent actions by PA700, the assembled 26 S proteasome displays greatly increased rates of energy-independent degradation of short peptides by virtue of their increased access to catalytic sites via diffusion through the open pore (15, 18, 20).Recently, specific interactions between Rpt and α subunits that determine PA700-20 S proteasome binding and gate opening have been defined. These findings established nonequivalent roles among the six different Rpt subunits for these processes (12, 19). For example, carboxypeptidase A treatment of PA700 selectively cleaves the C termini of two Rpt subunits (Rpt2 and Rpt5) and renders PA700 incompetent for proteasome binding and activation (19). Remarkably, short peptides corresponding to the C terminus of either Rpt2 or Rpt5, but none of the other Rpt subunits, were sufficient to bind to the 20 S proteasome and activate peptide substrate hydrolysis by inducing gate opening (12, 15, 18). The C-terminal peptides of Rpt2 and Rpt5 appear to bind to different and distinct sites on the proteasome and produce additive effects on rates of peptide substrate hydrolysis, suggesting that pore size or another feature of gating can be variably modulated (19). These various results, however, do not specify whether the action of one or the other or both C-terminal peptides is essential for function of intact PA700.In addition to its role in activation, PA700 plays other essential roles in 26 S proteasome function related to substrate selection and processing. For example, PA700 captures polyubiquitylated proteins via multiple subunits that bind polyubiquitin chains (2123). Moreover, to ensure translocation of the bound ubiquitylated protein through the narrow opened substrate access pore for proteolysis, PA700 destabilizes the tertiary structure of the protein via chaperone-like activity and removes polyubiquitin chains via deubiquitylating activities of several different subunits (2430). These various functions appear to be highly coordinated and may be mechanistically linked to one another and to the hydrolysis of ATP by Rpt subunits during substrate processing.Despite support for this general model of PA700 action, there is a lack of detailed knowledge about how PA700 subunits are structurally organized and functionally linked. Previously, we identified and characterized a subcomplex of PA700 called “modulator” that contained two ATPase subunits, Rpt4 and Rpt5, and one non-ATPase subunit, p27 (31). Although this protein was identified by an assay that measured increased PA700-dependent proteasome activation, the mechanistic basis of this effect was not clear. Moreover, the modulator lacked detectable ATPase activity and proteasome activating activity. The latter feature is surprising in retrospect because of the newly identified capacity of Rpt5 to activate the proteasome directly (12, 19). This disparity suggests that specific interactions among multiple PA700 subunits determine the manifestation and regulation of various activities.This study extends our recent findings regarding relative roles of Rpt subunits in the regulation of proteasome function. It also provides new insights and significance to older work that identified and characterized the modulator as a subcomplex of PA700. Our findings unite two different lines of investigation to offer new information about the structure, function, and regulation of 26 S proteasome. They also offer insights about alternative models for assembly of PA700 and 26 S proteasome in intact cells.  相似文献   

3.
The 26S proteasome is the end point of the ubiquitin-proteasome pathway and degrades ubiquitylated substrates. It is composed of the 20S core particle (CP), where degradation occurs, and the 19S regulatory particle (RP), which ensures substrate specificity of degradation. Whereas the CP is resolved to atomic resolution, the architecture of the RP is largely unknown. We provide a comprehensive analysis of the current structural knowledge on the RP, including structures of the RP subunits, physical protein-protein interactions, and cryoelectron microscopy data. These data allowed us to compute an atomic model for the CP-AAA-ATPase subcomplex. In addition to this atomic model, further subunits can be mapped approximately, which lets us hypothesize on the substrate path during its degradation.The ubiquitin-proteasome pathway is the major route used by eukaryotic cells for the disposal of misfolded or damaged proteins and for controlling the lifespan of proteins (13). As a consequence, the ubiquitin-proteasome pathway regulates a plethora of fundamental cellular processes, such as protein quality control, DNA repair, and signal transduction (4). The 26S proteasome is a molecular machine of ∼2.5 MDa that targets polyubiquitylated proteins. It comprises two subcomplexes, the 20S core particle (CP)1 and one or two copies of asymmetric 19S regulatory particles (RPs), which bind to the end(s) of the barrel-shaped CP.The active sites of the proteasome are located in the CP cavity where proteolytic cleavage of substrates takes place. Electron microscopy (EM) and x-ray crystallography have revealed that the CP is a cylinder consisting of four concentrically stacked rings (57): two identical “α”-rings, each assembled of seven homologous proteins, form the outer rings, and two identical “β”-rings, also assembled of seven homologs, form the two inner rings. Proteolysis is confined to the cavity formed by the β-rings, a nanocompartment sequestered from the cytosol.The RPs regulate substrate degradation by (i) binding polyubiquitylated substrates, (ii) subsequently deubiquitylating them, (iii) substrate unfolding, and (iv) opening the “gate” to the CP (8). The RPs consists of six AAA-ATPase subunits and at least 13 non-ATPase subunits. In contrast to the CP, the architecture of the RP subunits remains largely unknown. The problems that hamper structural characterization of the RP are manifold. It has proven difficult to obtain homogeneous, concentrated preparations of 26S proteasomes or RPs because the RP tends to disassociate into heterogeneous subcomplexes during purification and concentration. Moreover, many of the RP subunits likely exhibit a significant degree of structural variability. As a consequence, x-ray crystallographic analysis of the entire RP has not been accomplished to date, and only a few subunit fragments have been amenable to high resolution structure determination.For cryo-EM and protein-protein interaction experiments, the requirements for sample homogeneity are less stringent. Recently, the Drosophila melanogaster 26S proteasome was resolved to ∼20 Å (9). Various proteomics approaches have led to proposals for topological maps of the RP (1013). The resolution of protein-protein interaction data typically corresponds to the diameters of the proteins or domains found to interact, which are typically far beyond 20 Å. Because of the limited resolution, neither cryo-EM maps nor protein-protein interaction networks are by themselves sufficient to determine the RP architecture (i.e. the localization of the RP subunits in the complex).The integration of atomic models, cryo-EM maps, and protein-protein interaction data is currently the most promising approach to resolve the architecture of the 26S proteasome (1418). Here, we provide a comprehensive analysis of the current structural knowledge on the RP, including structures of RP subunits, physical protein-protein interactions, and cryo-EM data. Based on these data, we provide a model for the CP-AAA-ATPase subcomplex. Finally, we outline a path toward resolving the architecture of the 26S proteasome by an integrative structure determination approach, which in turn will provide a basis for a mechanistic understanding.  相似文献   

4.
5.
6.
7.
8.
Structural characterization of proteasome complexes is an essential step toward understanding the ubiquitin-proteasome system. Currently, high resolution structures are not available for the 26S proteasome holocomplex as well as its subcomplex, the 19S regulatory particle (RP). Here we have employed a novel integrated strategy combining chemical cross-linking with multistage tandem mass spectrometry to define the proximity of subunits within the yeast 19S RP to elucidate its topology. This has resulted in the identification of 174 cross-linked peptides of the yeast 19S RP, representing 43 unique lysine-lysine linkages within 24 nonredundant pair-wise subunit interactions. To map the spatial organization of the 19S RP, we have developed and utilized a rigorous probabilistic framework to derive maximum likelihood (ML) topologies based on cross-linked peptides determined from our analysis. Probabilistic modeling of the yeast 19S AAA-ATPase ring (i.e., Rpt1–6) has produced an ML topology that is in excellent agreement with known topologies of its orthologs. In addition, similar analysis was carried out on the 19S lid subcomplex, whose predicted ML topology corroborates recently reported electron microscopy studies. Together, we have demonstrated the effectiveness and potential of probabilistic modeling for unraveling topologies of protein complexes using cross-linking data. This report describes the first study of the 19S RP topology using a new integrated strategy combining chemical cross-linking, mass spectrometry, and probabilistic modeling. Our results have provided a solid foundation to advance our understanding of the 19S RP architecture at peptide level resolution. Furthermore, our methodology developed here is a valuable proteomic tool that can be generalized for elucidating the structures of protein complexes.Basic cellular homeostasis depends on the regulated protein degradation and turnover by the ubiquitin-proteasome system (1, 2). Central to this pathway is the 26S proteasome complex, which is responsible for ubiquitin/ATP-dependent protein degradation (35). The 26S holocomplex is a megadalton-sized protein assembly consisting of the 20S catalytic core particle (CP)1 and the 19S regulatory particle (RP). The eukaryotic 20S CP is composed of two copies of 14 nonidentical subunits (α1–7 and β1–7) arranged into four stacked heptameric rings in an order of α7β7β7α7. The crystal structure and topology of the highly ordered 20S CP has been resolved and is evolutionarily conserved (6). Although α subunits of the 20S CP are essential for the assembly of the complex and its interactions with the regulatory complex, three catalytic β subunits (β1, β2, and β5) harbor various catalytic activities responsible for regulated proteasomal degradation. The 19S RP is composed of 19 subunits, which forms two subcomplexes, the base consisting of six related AAA-ATPase (Rpt1–6) and four non-ATPase (Rpn1, Rpn2, Rpn10, and Rpn13) subunits and the lid containing nine non-ATPase subunits (Rpn3, Rpn5–9, Rpn11, Rpn12, and Rpn15/Sem1) (7, 8). In comparison with the 20S core, the function and structure of the 19S RP is much less well understood. Nevertheless, it is believed that the 19S RP is involved in multiple functions including recognition of polyubiquitinated substrates (9, 10), cleavage of the polyubiquitin chains to recycle ubiquitin (11), unfolding of substrates, assisting in opening the gate of the 20S chamber, and subsequently translocating the unfolded substrates into the catalytic chamber (4, 1214). The six AAA-ATPase subunits (Rpt1–6), which directly interact with the 20S α-ring, function as a molecular chaperone responsible for protein unfolding and are involved in substrate translocation and modulating gating of the CP (5, 15). Although detailed functions for most of the 19S non-ATPase subunits remain elusive, Rpn11 is known to carry an Mpr1p and Pad1p N-termini (MPN) domain, which harbors an essential deubiquitination activity responsible for cleaving polyubiquitin chains from proteasomal substrates (11, 16). In addition, two proteasome subunits, Rpn10 and Rpn13, have been identified as ubiquitin receptors, which are important in docking ubiquitinated substrates to the proteasome for degradation (4). Moreover, the two largest proteasome subunits, Rpn1 and Rpn2, interact with a variety of proteins including ubiquitin receptors and deubiquitinases and thus may function as scaffolding proteins to assist proteasomal degradation. Thus far, no atomic resolution structures are available for either the 19S RP or the 26S holocomplex. New insights of the overall topology of the 19S RP will illuminate protein interactions within, thus providing evidence for its otherwise unknown functions.Although many studies have been performed to characterize the 19S structure utilizing various techniques including cryo-EM (17, 18) and native mass spectrometry (19), details on spatial interfaces and subunit interconnectivity of the 19S RP remain to be unraveled. During the course of our study, the rough topology of the 19S RP was determined by cryo-EM alone (20) or coupled with other approaches (21); nevertheless more detailed information at the peptide or atomic level is still required. In addition to technological limitations in current approaches, the highly dynamic and heterogeneous nature of the 19S RP may attribute to the difficulty in obtaining its high resolution structure. In recent years, chemical cross-linking coupled with mass spectrometry (XL-MS) has become an attractive alternative for structure analysis of proteins and protein complexes (22, 23). The ability of XL-MS to identify interaction interfaces between proteins allows us to define low resolution protein topology. In addition to protein interaction networks and the site of protein interactions at binding interfaces, cross-linking analysis can reveal information about the spatial distance between cross-linked amino acids on the surface of folded proteins. Although such knowledge only reveals the maximum distance given by the length of the cross-linker and can be influenced by protein conformational flexibility, it can be used as the distance constraint for molecular modeling of protein folds and complex topologies, i.e., the arrangement of the constituents of a complex in space. A recent study by Chen et al. (24) on yeast RNA polymerase II (RNAPII) complex has exemplified the power of XL-MS in elucidating the architecture of large multisubunit complexes. Although effective, cross-linking studies have been challenging because of the low abundance of cross-linked products and the inherent complexity of sequencing interlinked peptides by MS for unambiguous identification. To facilitate MS detection and identification of cross-linked products, we have recently developed a novel homobifunctional amine reactive, low energy MS-cleavable cross-linker, disuccinimidyl sulfoxide (DSSO), and successfully applied it to cross-link the yeast 20S proteasome for rapid, accurate, and simplified determination of protein interaction interfaces within the complex (25). The unique functionality of our cross-linking reagent and specialized bioinformatics tools significantly increase our confidence and speed in the identification of cross-linked products when compared with cross-linking studies using traditional noncleavable reagents. Current cross-linking studies have been focused on protein complexes with known crystal structures, but topological structures of protein complexes based primarily on cross-linking data have not yet been reported. This is due to the lack of computational tools that use cross-linking data to deduce the spatial organization of subunits in a given complex. To define the architecture of the yeast 19S RP, we have characterized the proximity and interconnectivity of the subunits by employing our newly developed cross-linking strategy. The resulting cross-linking information serves as a basis for a rigorous probabilistic analysis to obtain the maximum likelihood (ML) topology. This strategy is developed by first analyzing our cross-linking data for the 19S six-member AAA-ATPase base ring, as the topology ordering of yeast orthologs has been recently determined (14, 2628). The effectiveness of this new probabilistic platform is supported by the agreement between our derived ML topology of the AAA-ATPase base ring and previous reports. When the same probabilistic approach is applied to the 19S lid subcomplex, the resulting topology is also in agreement with recently proposed models (20, 21). This work represents the first application of probabilistic modeling of protein complexes based solely on cross-link data, establishing a new workflow for future structural analysis of large protein complexes using XL-MS.  相似文献   

9.
10.
The impact of Nα-terminal acetylation on protein stability and protein function in general recently acquired renewed and increasing attention. Although the substrate specificity profile of the conserved enzymes responsible for Nα-terminal acetylation in yeast has been well documented, the lack of higher eukaryotic models has hampered the specificity profile determination of Nα-acetyltransferases (NATs) of higher eukaryotes. The fact that several types of protein N termini are acetylated by so far unknown NATs stresses the importance of developing tools for analyzing NAT specificities. Here, we report on a method that implies the use of natural, proteome-derived modified peptide libraries, which, when used in combination with two strong cation exchange separation steps, allows for the delineation of the in vitro specificity profiles of NATs. The human NatA complex, composed of the auxiliary hNaa15p (NATH/hNat1) subunit and the catalytic hNaa10p (hArd1) and hNaa50p (hNat5) subunits, cotranslationally acetylates protein N termini initiating with Ser, Ala, Thr, Val, and Gly following the removal of the initial Met. In our studies, purified hNaa50p preferred Met-Xaa starting N termini (Xaa mainly being a hydrophobic amino acid) in agreement with previous data. Surprisingly, purified hNaa10p preferred acidic N termini, representing a group of in vivo acetylated proteins for which there are currently no NAT(s) identified. The most prominent representatives of the group of acidic N termini are γ- and β-actin. Indeed, by using an independent quantitative assay, hNaa10p strongly acetylated peptides representing the N termini of both γ- and β-actin, and only to a lesser extent, its previously characterized substrate motifs. The immunoprecipitated NatA complex also acetylated the actin N termini efficiently, though displaying a strong shift in specificity toward its known Ser-starting type of substrates. Thus, complex formation of NatA might alter the substrate specificity profile as compared with its isolated catalytic subunits, and, furthermore, NatA or hNaa10p may function as a post-translational actin Nα-acetyltransferase.The multisubunit and ribosome-associated protein Nα-acetyltransferases (NATs)1 are omnipresent enzyme complexes that catalyze the transfer of the acetyl moiety from acetyl-CoA to the primary α-amines of N termini of nascent proteins (13). As up to 50 to 60% of yeast proteins and 80 to 90% of human proteins are modified in this manner, Nα-acetylation is a widespread protein modification in eukaryotes (47), and the pattern of modification has remained largely conserved throughout evolution (4, 8). NATs belong to a subfamily of the Gcn5-related N-acetyltransferase superfamily of N-acetyltransferases, additionally encompassing the well-studied histone acetyltransferases that are implicated in epigenetic imprinting.In yeast and humans, three main NAT complexes, NatA, NatB, and NatC were found to be responsible for the majority of Nα-terminal acetylations (1). The NatA complex, responsible for cotranslational Nα-terminal acetylation of proteins with Ser, Ala, Thr, Gly, and Val N termini, is composed of two main subunits, the catalytic subunit Naa10p (previously known as Ard1p) and the auxiliary subunit Naa15p (previously known as Nat1p/NATH) (911). Furthermore, a third catalytic subunit Naa50p (previously known as Nat5)—an acetyltransferase shown to function in chromosome cohesion and segregation (1214)—was found to physically interact with the NatA complex of yeast (2), fruit fly (12), and human (15). Recently, human Naa50p (hNaa50p) was reported to display lysine or Nε-acetyltransferase as well as NAT activity (16), the latter was defined as NatE activity (16). Interestingly, the chaperone-like, Huntingtin interacting protein HYPK, identified as a novel stable interactor of human NatA, was functionally implicated in the N-terminal acetylation of an in vivo NatA substrate, demonstrating that NAT complex formation and composition may have an overall influence on the observed (degree of) Nα-acetylation (17). Further, subunits of the human NatA complex have been coupled to cancer-related processes and differentiation, with altered subunit expression reported in papillary thyroid carcinoma, neuroblastoma, and retinoic acid induced differentiation. Furthermore, the NatA catalytic subunit was found to be implicated in processes such as hypoxia-response and the β-catenin pathway (18, 19). Of note is that in line with the differential localization patterns of the individual NatA subunits (9, 13, 20, 21), other data indicate that these subunits might well exert NatA-independent enzymatic functions (13, 22, 23). Given that a significant fraction of hNaa10p and hNaa15p are nonribosomal (9), and given the multitude of postulated post-translational in vivo N-acetylation events recently reported (2426), these observations argue in favor of the existence of NAT complexes and/or catalytic NAT-subunits acting post-translationally.Similar to NatA, the NatB and NatC complexes, composed of the catalytic subunit Naa20p or Naa30p and the auxiliary subunits Naa25p or Naa35p and Naa38p respectively, are conserved from yeast to higher eukaryotes concerning their subunit composition as well as their substrate specificity. Both these complexes display activity toward methionine-starting N termini, with NatB preferring acidic residues as well as Asn and Gln at P2′-sites2, whereas NatC prefers hydrophobic amino acid residues at substrate P2′-sites (1, 27, 28).Nα-acetylation affects various protein functions such as localization, activity, association, and stability (29, 30). Only recently a more generalized function of protein Nα-acetylation in generating so-called N-terminal degrons marking proteins for removal was put forward (31). The lack of mouse models in addition to the fact that (combined) knockdown of individual components of Nα-acetyltransferases only marginally affect the overall Nα-acetylation status (4) have so far hampered the molecular characterization of the substrate specificity profile of (yet uncharacterized) NATs. To date, all eukaryote Nα-acetylation events are assumed to be catalyzed by the five known NATs (32). However, an additional level of complexity is imposed by the fact that in contrast to yeast, higher eukaryotes express multiple splice variants of various NAT subunits as well as paralogs thereof (33, 34), further implicating that a specific NAT''s substrate specificity might be altered in this way, in addition to the possible existence of substrate redundancy. Moreover, regulation of substrate specificity and stability of NAT activity can be imposed by differential complex formation and post-translational modifications including phosphorylation, auto-acetylation, and specific proteolytic cleavage of the catalytic subunits (9, 16, 17). As such, a detailed understanding of the substrate specificity of NATs, and the regulation thereof, could help unravel the physiological substrate repertoires as well as the associated physiological roles of NATs in the normal and the disease state.The specificity of Nα-acetyltransferases and their endogenous substrates were originally studied by two-dimensional-PAGE: Nα-acetylation neutralizes the N-terminal positive charge, resulting in an altered electrophoretic protein migration during isoelectric focusing (3538). Recently, this altered biophysical property was also exploited to enrich for protein N-termini using low pH strong cation exchange (SCX) chromatography (24, 39). As an example, SCX prefractionation combined with N-terminal combined fractional diagonal chromatography, a targeted proteomics technology negatively selecting for protein N-terminal peptides, stable isotope labeling of amino acids in cell culture, and amino-directed modifiers (40), was used to study the in vivo substrate repertoires of human as well as yeast NatA (4).Nevertheless, the various methods reported today to study in detail Nα-terminal acetylation and thus the specificities of different NATs make use of a limited and therefore somewhat biased set of synthesized peptide substrates and comprise the rather laborious detection of radioactive acetylated products as well as enzyme-coupled methods quantifying acetyl-CoA conversion. Because (proteome-derived) peptide libraries have been used extensively to study epitope mapping (41), protein-protein interactions (42), protein modifications such as phosphorylation (43), and proteolysis (44, 45), as well as for determining the substrate specificity of the Nα-deblocking peptide deformylase (46), we reckoned that the development of an oligopeptide-based acetylation assay should allow for more comprehensive screening of NAT-like activities. We here report on the development of a peptide-based method to systematically screen for the in vitro sequence specificity profile of individual NATs as well as endogenous NAT complexes. In summary, SCX enriched, Nα-free peptide libraries, derived from natural proteomes build up the peptide substrate pool. And, upon incubation, NAT Nα-acetylated peptides are enriched by a second SCX fractionation step, resulting in a positive selection of NAT-specific peptide substrates. By use of this proteome-derived peptide library approach, we here delineated (differences in) the specificity profiles of hNaa50p and hNaa10p as isolated hNatA components, as well as of assayed their combined activity when in their native hNatA complex.  相似文献   

11.
12.
13.
14.
15.
16.
Congenital muscular dystrophy with laminin α2 chain deficiency (MDC1A) is one of the most severe forms of muscular disease and is characterized by severe muscle weakness and delayed motor milestones. The genetic basis of MDC1A is well known, yet the secondary mechanisms ultimately leading to muscle degeneration and subsequent connective tissue infiltration are not fully understood. In order to obtain new insights into the molecular mechanisms underlying MDC1A, we performed a comparative proteomic analysis of affected muscles (diaphragm and gastrocnemius) from laminin α2 chain–deficient dy3K/dy3K mice, using multidimensional protein identification technology combined with tandem mass tags. Out of the approximately 700 identified proteins, 113 and 101 proteins, respectively, were differentially expressed in the diseased gastrocnemius and diaphragm muscles compared with normal muscles. A large portion of these proteins are involved in different metabolic processes, bind calcium, or are expressed in the extracellular matrix. Our findings suggest that metabolic alterations and calcium dysregulation could be novel mechanisms that underlie MDC1A and might be targets that should be explored for therapy. Also, detailed knowledge of the composition of fibrotic tissue, rich in extracellular matrix proteins, in laminin α2 chain–deficient muscle might help in the design of future anti-fibrotic treatments. All MS data have been deposited in the ProteomeXchange with identifier PXD000978 (http://proteomecentral.proteomexchange.org/dataset/PXD000978).Congenital muscular dystrophy with laminin α2 chain deficiency, also known as MDC1A,1 is a severe muscle wasting disease for which there is no cure. MDC1A is caused by mutations in the LAMA2 gene that lead to complete or partial deficiency of laminin α2 chain (13). Although the primary defect in MDC1A is known, the secondary molecular mechanisms eventually leading to muscle degeneration are not fully understood. In normal muscle, laminin α2 chain binds to the cell surface receptors dystroglycan and integrin α7β1, which both indirectly bind the cytoskeleton (47). Both of these adhesion complexes are important for normal skeletal muscle function, and laminin α2 chain binding to dystroglycan contributes to the maintenance of sarcolemmal integrity and protects muscles from damage (8), whereas laminin α2 chain binding to integrin α7β1 promotes myofiber survival (9, 10). In MDC1A, laminin α2 chain is absent or severely reduced, and the expression of dystroglycan and α7β1 is also dysregulated in MDC1A (9, 11, 12). Thus, the structural link is broken, and the yet to be determined downstream intracellular signaling pathways are also interrupted. Consequently, laminin α2 chain–deficient muscle fibers undergo degeneration–regeneration cycles, but rather quickly regeneration fails and muscle fibers die by apoptosis/necrosis followed by a major replacement of muscle tissue with connective tissue (3, 7). In order to unravel novel secondary molecular mechanisms, which could indicate new therapeutic targets, we decided to evaluate the protein expression profile in laminin α2 chain–deficient dy3K/dy3K muscle. Several proteomic profiling studies of dystrophin-deficient muscles (Duchenne muscular dystrophy) have been performed (1320), as well as some with dysferlin-deficient muscles (Limb-girdle muscular dystrophy type 2B, Miyoshi myopathy) (21, 22). They all showed a great number of proteins that were differentially expressed in different dystrophic muscles and at different ages (1322). However, proteomic analyses of laminin α2 chain–deficient muscle have not yet been performed. We here used multidimensional protein identification technology with tandem mass tags (TMT), a powerful shotgun label-based proteomic method that separates peptides in two-dimensional liquid chromatography (23, 24). We identified around 100 proteins that were differentially expressed in laminin α2 chain–deficient gastrocnemius and diaphragm muscles relative to the corresponding wild-type muscles, and the differential expression of selected proteins was verified with Western blot analysis or immunofluorescence.  相似文献   

17.
18.
19.
Detection of endogenous ubiquitination sites by mass spectrometry has dramatically improved with the commercialization of anti-di-glycine remnant (K-ε-GG) antibodies. Here, we describe a number of improvements to the K-ε-GG enrichment workflow, including optimized antibody and peptide input requirements, antibody cross-linking, and improved off-line fractionation prior to enrichment. This refined and practical workflow enables routine identification and quantification of ∼20,000 distinct endogenous ubiquitination sites in a single SILAC experiment using moderate amounts of protein input.The commercialization of antibodies that recognize lysine residues modified with a di-glycine remnant (K-ε-GG)1 has significantly transformed the detection of endogenous protein ubiquitination sites by mass spectrometry (15). Prior to the development of these highly specific reagents, proteomics experiments were limited to identification of up to only several hundred ubiquitination sites, which severely limited the scope of global ubiquitination studies (6). Recent proteomic studies employing anti-K-ε-GG antibodies have enhanced our understanding of ubiquitin biology through the identification of thousands of ubiquitination sites and the analysis of the change in relative abundance of these sites after chemical or biological perturbation (13, 5, 7). Use of stable isotope labeling by amino acids in cell culture (SILAC) for quantification has enabled researchers to better understand the extent of ubiquitin regulation upon proteasome inhibition and precisely identify those protein classes, such as newly synthesized proteins or chromatin-related proteins, that see overt changes in their ubiquitination levels upon drug treatment (2, 3, 5). Emanuel et al. (1) have combined genetic and proteomics assays implementing the anti-K-ε-GG antibody to identify hundreds of known and putative Cullin-RING ligase substrates, which has clearly demonstrated the extensive role of Cullin-RING ligase ubiquitination on cellular protein regulation.Despite the successes recently achieved with the use of the anti-K-ε-GG antibody, increased sample input (up to ∼35 mg) and/or the completion of numerous experimental replicates have been necessary to achieve large numbers of K-ε-GG sites (>5,000) in a single SILAC-based experiment (13, 5). For example, it has been recently shown that detection of more than 20,000 unique ubiquitination sites is possible from the analysis of five different murine tissues (8). However, as the authors indicate, only a few thousands sites are detected in any single analysis of an individual tissue sample (8). It is recognized that there is need for further improvements in global ubiquitin technology to increase the depth-of-coverage attainable in quantitative proteomic experiments using moderate amounts of protein input (9). Through systematic study and optimization of key pre-analytical variables in the preparation and use of the anti-K-ε-GG antibody as well as the proteomic workflow, we have now achieved, for the first time, routine quantification of ∼20,000 nonredundant K-ε-GG sites in a single SILAC triple encoded experiment starting with 5 mg of protein per SILAC channel. This represents a 10-fold improvement over our previously published method (3).  相似文献   

20.
The human JC polyomavirus (JCV) is the etiologic agent of the fatal central nervous system (CNS) demyelinating disease progressive multifocal leukoencephalopathy (PML). PML typically occurs in immunosuppressed patients and is the direct result of JCV infection of oligodendrocytes. The initial event in infection of cells by JCV is attachment of the virus to receptors present on the surface of a susceptible cell. Our laboratory has been studying this critical event in the life cycle of JCV, and we have found that JCV binds to a limited number of cell surface receptors on human glial cells that are not shared by the related polyomavirus simian virus 40 (C. K. Liu, A. P. Hope, and W. J. Atwood, J. Neurovirol. 4:49–58, 1998). To further characterize specific JCV receptors on human glial cells, we tested specific neuraminidases, proteases, and phospholipases for the ability to inhibit JCV binding to and infection of glial cells. Several of the enzymes tested were capable of inhibiting virus binding to cells, but only neuraminidase was capable of inhibiting infection. The ability of neuraminidase to inhibit infection correlated with its ability to remove both α(2-3)- and α(2-6)-linked sialic acids from glial cells. A recombinant neuraminidase that specifically removes the α(2-3) linkage of sialic acid had no effect on virus binding or infection. A competition assay between virus and sialic acid-specific lectins that recognize either the α(2-3) or the α(2-6) linkage revealed that JCV preferentially interacts with α(2-6)-linked sialic acids on glial cells. Treatment of glial cells with tunicamycin, but not with benzyl N-acetyl-α-d-galactosaminide, inhibited infection by JCV, indicating that the sialylated JCV receptor is an N-linked glycoprotein. As sialic acid containing glycoproteins play a fundamental role in mediating many virus-cell and cell-cell recognition processes, it will be of interest to determine what role these receptors play in the pathogenesis of PML.Approximately 70% of the human population worldwide is seropositive for JC virus (JCV). Like other polyomaviruses, JCV establishes a lifelong latent or persistent infection in its natural host (40, 49, 50, 68, 72). Reactivation of JCV in the setting of an underlying immunosuppressive illness, such as AIDS, is thought to lead to virus dissemination to the central nervous system (CNS) and subsequent infection of oligodendrocytes (37, 40, 66, 68). Reactivation of latent JCV genomes already present in the CNS has also been postulated to contribute to the development of progressive multifocal leukoencephalopathy (PML) following immunosuppression (19, 48, 55, 70, 75). Approximately 4 to 6% of AIDS patients will develop PML during the course of their illness (10). In the CNS, JCV specifically infects oligodendrocytes and astrocytes. Outside the CNS, JCV genomes have been identified in the urogenital system, in the lymphoid system, and in B lymphocytes (2, 17, 18, 30, 47, 59). In vitro, JCV infects human glial cells and, to a limited extent, human B lymphocytes (3, 4, 39, 41, 42). Recently, JCV infection of tonsillar stromal cells and CD34+ B-cell precursors has been described (47). These observations have led to the suggestion that JCV may persist in a lymphoid compartment and that B cells may play a role in trafficking of JCV to the CNS (4, 30, 47).Virus-receptor interactions play a major role in determining virus tropism and tissue-specific pathology associated with virus infection. Viruses that have a very narrow host range and tissue tropism, such as JCV, are often shown to interact with high affinity to a limited number of specific receptors present on susceptible cells (26, 44). In some instances, virus tropism is strictly determined by the presence of specific receptors that mediate binding and entry (7, 16, 27, 35, 46, 53, 56, 67, 73, 74, 76). In other instances, however, successful entry into a cell is necessary but not sufficient for virus growth (5, 8, 45, 57). In these cases, additional permissive factors that interact with viral regulatory elements are required.The receptor binding characteristics of several polyomaviruses have been described. The mouse polyomavirus (PyV) receptor is an N-linked glycoprotein containing terminal α(2-3)-linked sialic acid (1214, 22, 28). Both the large and small plaque strains of PyV recognize α(2-3)-linked sialic acid. The small-plaque strain also recognizes a branched disialyl structure containing α(2-3)- and α(2-6)-linked sialic acids. Neither strain recognizes straight-chain α(2-6)-linked sialic acid. The ability of the large- and small-plaque strains of PyV to differentially recognize these sialic acid structures has been precisely mapped to a single amino acid in the major virus capsid protein VP1 (21). The large-plaque strains all contain a glycine at amino acid position 92 in VP1, and the small-plaque strains all contain a negatively charged glutamic acid at this position (21). In addition to forming small or large plaques, these strains also differ in the ability to induce tumors in mice (20). This finding suggests that receptor recognition plays an important role in the pathogenesis of PyV.The cell surface receptor for lymphotropic papovavirus (LPV) is an O-linked glycoprotein containing terminal α(2-6)-linked sialic acid (26, 33, 34). Infection with LPV is restricted to a subset of human B-cell lines, and recognition of specific receptors is a major determinant of the tropism of LPV for these cells (26).Unlike the other members of the polyomavirus family, infection of cells by simian virus 40 (SV40) is independent of cell surface sialic acids. Instead, SV40 infection is mediated by major histocompatibility complex (MHC)-encoded class I proteins (5, 11). MHC class I proteins also play a role in mediating the association of SV40 with caveolae, a prerequisite for successful targeting of the SV40 genome to the nucleus of a cell (1, 63). Not surprisingly, SV40 has been shown not to compete with the sialic acid-dependent polyomaviruses for binding to host cells (15, 26, 38, 58).Very little is known about the early steps of JCV binding to and infection of glial cells. Like other members of the polyomavirus family, JCV is known to interact with cell surface sialic acids (51, 52). A role for sialic acids in mediating infection of glial cells has not been described. It is also not known whether the sialic acid is linked to a glycoprotein or a glycolipid. In a previous report, we demonstrated that JCV bound to a limited number of cell surface receptors on SVG cells that were not shared by the related polyomavirus SV40 (38). In this report, we demonstrate that virus binding to and infection of SVG cells is dependent on an N-linked glycoprotein containing terminal α(2-3)- and α(2-6)-linked sialic acids. Competitive binding assays with sialic acid-specific lectins suggest that the virus preferentially interacts with α(2-6)-linked sialic acids. We are currently evaluating the role of this receptor in determining the tropism of JCV for glial cells and B cells.  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