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.
The ubiquitin-proteasome system is a central mechanism for controlled proteolysis that regulates numerous cellular processes in eukaryotes. As such, defects in this system can contribute to disease pathogenesis. In this pathway, E3 ubiquitin ligases provide platforms for binding specific substrates, thereby coordinating their ubiquitylation and subsequent degradation by the proteasome. Despite the identification of many E3 ubiquitin ligases, the identities of their specific substrates are still largely unresolved. The ankyrin repeat-containing protein with a suppressor of cytokine signaling box 2 (ASB2) gene that we initially identified as a retinoic acid-response gene in acute promyelocytic leukemia cells encodes the specificity subunit of an E3 ubiquitin ligase complex that is involved in hematopoietic cell differentiation. We have recently identified filamin A and filamin B as the first ASB2 targets and shown that ASB2 triggers ubiquitylation and proteasome-mediated degradation of these proteins. Here a global quantitative proteomics strategy is provided to identify substrates of E3 ubiquitin ligases targeted to proteasomal degradation. Indeed we used label-free methods for quantifying proteins identified by shotgun proteomics in extracts of cells expressing wild-type ASB2 or an E3 ubiquitin ligase-defective mutant of ASB2 under the control of an inducible promoter. Measurements of spectral count and mass spectrometric signal intensity demonstrated a drastic decrease of filamin A and filamin B in myeloid leukemia cells expressing wild-type ASB2 compared with cells expressing an E3 ubiquitin ligase-defective mutant of ASB2. Altogether we provide an original strategy that enables identification of E3 ubiquitin ligase substrates that have to be degraded.The ubiquitin-proteasome system (UPS)1 plays an essential role in the regulation of protein stability in eukaryotic cells. Degradation of a protein by the UPS entails two successive steps: the covalent attachment of multiple ubiquitin molecules to the protein substrate and its degradation by the 26 S proteasome (1, 2). Ubiquitylation of protein substrates occurs through the sequential action of distinct enzymes: a ubiquitin-activating enzyme, E1; a ubiquitin-conjugating enzyme, E2; and a ubiquitin ligase, E3, responsible for the specific recognition of substrates. Increasing attention has been recently given to the UPS leading to the identification of hundreds of E3 ubiquitin ligases (E3s). Two major classes of E3s have been described: (i) E3s of the HECT (homologous to the E6-associated protein carboxyl terminus) domain family that function as ubiquitin carriers (3, 4) and (ii) E3s of the RING (really interesting new gene) or of the U box families that have no inherent catalytic activity but recruit an E2 enzyme toward substrates (57).Classical approaches to identify substrates of E3s are based on the identification of interacting proteins. Although these have successfully led to the identification of a number of substrates of monomeric E3s, identification of substrates of multimeric E3s is very challenging because of the weak affinity of substrates for their requisite specificity subunit and because of the labile nature of the substrate complexed with the specificity subunit (8).Acute promyelocytic leukemia (APL) is associated with six reciprocal translocations always involving the retinoic acid receptor α (RARα) gene (911). The RARα protein is a member of the nuclear receptor superfamily that stimulates myeloid differentiation in the presence of its ligand, all-trans-retinoic acid (RA). In more than 95% of APL, the t(15;17) translocation between the promyelocytic leukemia (PML) gene on chromosome 15 and the RARα gene on chromosome 17 produces the PML-RARα fusion protein (12). The PML-RARα protein enhances the repression of RARα target genes by increasing associations with corepressors (1315) and by recruiting DNA methyltransferases (16). These complexes dissociate from the PML-RARα fusion protein in the presence of pharmacological concentrations of RA perhaps explaining why APL cells are sensitive to RA treatment. Indeed at pharmacological concentrations, RA induces complete remission in a high percentage of APL patients (1719). By studying RA-induced differentiation of APL cells we have attempted to identify some of the genes that may be up-regulated during this process to further understand the control of growth and differentiation in leukemia (20). One gene identified in this manner, ASB2 (ankyrin repeat-containing protein with a suppressor of cytokine signaling box 2) is an RA-response gene involved in induced differentiation of myeloid leukemia cells (2123).The ASB2 protein is a subunit of a multimeric E3 ubiquitin ligase of the cullin-RING ligase family (24, 25). The ASB2 suppressor of cytokine signaling box can be divided into a BC box that defines a binding site for the Elongin BC complex and a Cul5 box that determines the binding specificity for Cullin5 (24, 26). Indeed the ASB2 protein, by interacting with the Elongin BC complex, can assemble with a Cullin5/Rbx1 or -2 module to reconstitute an active E3 ubiquitin ligase complex (2325). Within this complex, the ASB2 protein is the specificity subunit involved in the recruitment of specific substrate(s). Furthermore endogenous ASB2 protein was copurified with ubiquitin ligase activity in RA-treated APL cells suggesting that, during induced differentiation of leukemia cells, the ASB2 protein may target proteins involved in blocking differentiation to destruction by the proteasome machinery (24). We recently identified actin-binding proteins filamin A (FLNa) and filamin B (FLNb) as ASB2 targets and showed that ASB2 triggers ubiquitylation and drives proteasome-mediated degradation of these proteins during RA-induced differentiation of myeloid leukemia cells (23).With the aim to develop a strategy to identify E3 substrates that are degraded by the proteasome, we used an MS approach to identify ASB2 substrates in physiologically relevant settings. Indeed we used label-free quantitative proteomics to identify proteins that are absent or less abundant in cells that express wild-type ASB2 but that accumulate in cells expressing an ASB2 E3 ligase-defective mutant. Application of label-free MS methods that have the advantage to be simple, fast, and cheap enabled the identification of FLNa and FLNb as ASB2 substrates. This study provides a new strategy for the identification of E3 substrates that have to be degraded.  相似文献   

3.
4.
The human immunodeficiency virus type 1 (HIV-1) vpu gene encodes a type I anchored integral membrane phosphoprotein with two independent functions. First, it regulates virus release from a post-endoplasmic reticulum (ER) compartment by an ion channel activity mediated by its transmembrane anchor. Second, it induces the selective down regulation of host cell receptor proteins (CD4 and major histocompatibility complex class I molecules) in a process involving its phosphorylated cytoplasmic tail. In the present work, we show that the Vpu-induced proteolysis of nascent CD4 can be completely blocked by peptide aldehydes that act as competitive inhibitors of proteasome function and also by lactacystin, which blocks proteasome activity by covalently binding to the catalytic β subunits of proteasomes. The sensitivity of Vpu-induced CD4 degradation to proteasome inhibitors paralleled the inhibition of proteasome degradation of a model ubiquitinated substrate. Characterization of CD4-associated oligosaccharides indicated that CD4 rescued from Vpu-induced degradation by proteasome inhibitors is exported from the ER to the Golgi complex. This finding suggests that retranslocation of CD4 from the ER to the cytosol may be coupled to its proteasomal degradation. CD4 degradation mediated by Vpu does not require the ER chaperone calnexin and is dependent on an intact ubiquitin-conjugating system. This was demonstrated by inhibition of CD4 degradation (i) in cells expressing a thermally inactivated form of the ubiquitin-activating enzyme E1 or (ii) following expression of a mutant form of ubiquitin (Lys48 mutated to Arg48) known to compromise ubiquitin targeting by interfering with the formation of polyubiquitin complexes. CD4 degradation was also prevented by altering the four Lys residues in its cytosolic domain to Arg, suggesting a role for ubiquitination of one or more of these residues in the process of degradation. The results clearly demonstrate a role for the cytosolic ubiquitin-proteasome pathway in the process of Vpu-induced CD4 degradation. In contrast to other viral proteins (human cytomegalovirus US2 and US11), however, whose translocation of host ER molecules into the cytosol occurs in the presence of proteasome inhibitors, Vpu-targeted CD4 remains in the ER in a transport-competent form when proteasome activity is blocked.

The human immunodeficiency virus type 1 (HIV-1)-specific accessory protein Vpu performs two distinct functions in the viral life cycle (11, 12, 29, 34, 46, 47, 5052; reviewed in references 31 and 55): enhancement of virus particle release from the cell surface, and the selective induction of proteolysis of newly synthesized membrane proteins. Known targets for Vpu include the primary virus receptor CD4 (63, 64) and major histocompatibility complex (MHC) class I molecules (28). Vpu is an oligomeric class I integral membrane phosphoprotein (35, 48, 49) with a structurally and functionally defined domain architecture: an N-terminal transmembrane anchor and C-terminal cytoplasmic tail (20, 34, 45, 47, 50, 65). Vpu-induced degradation of endoplasmic reticulum (ER) membrane proteins involves the phosphorylated cytoplasmic tail of the protein (50), whereas the virion release function is mediated by a cation-selective ion channel activity associated with the membrane anchor (19, 31, 45, 47).CD4 is a 55-kDa class I integral membrane glycoprotein that serves as the primary coreceptor for HIV entry into cells. CD4 consists of a large lumenal domain, a transmembrane peptide, and a 38-residue cytoplasmic tail. It is expressed on the surface of a subset of T lymphocytes that recognize MHC class II-associated peptides, and it plays a pivotal role in the development and maintenance of the immune system (reviewed in reference 30). Down regulation of CD4 in HIV-1-infected cells is mediated through several independent mechanisms (reviewed in references 5 and 55): intracellular complex formation of CD4 with the HIV envelope protein gp160 (8, 14), endocytosis of cell surface CD4 induced by the HIV-1 nef gene product (1, 2), and ER degradation induced by the HIV-1 vpu gene product (63, 64).Vpu-induced degradation of CD4 is an example of ER-associated protein degradation (ERAD). ERAD is a common outcome when proteins in the secretory pathway are unable to acquire their native structure (4). Although it was thought that ERAD occurs exclusively inside membrane vesicles of the ER or other related secretory compartments, this has gained little direct experimental support. Indeed, there are several recent reports that ERAD may actually represent export of the target protein to the cytosol, where it is degraded by cytosolic proteases. It was found that in yeast, a secreted protein, prepro-α-factor (pαF), is exported from microsomes and degraded in the cytosol in a proteasome-dependent manner (36). This process was dependent on the presence of calnexin, an ER-resident molecular chaperone that interacts with N-linked oligosaccharides containing terminal glucose residues (3). In mammalian cells, two human cytomegalovirus (HCMV) proteins, US2 and US11, were found to cause the retranslocation of MHC class I molecules from the ER to the cytosol, where they are destroyed by proteasomes (61, 62). In the case of US2, class I molecules were found to associate with a protein (Sec61) present in the channel normally used to translocate newly synthesized proteins into the ER (termed the translocon), leading to the suggestion that the ERAD substrates are delivered to the cytosol by retrograde transport through the Sec61-containing pore (61). Fujita et al. (24) reported that, similar to these findings, the proteasome-specific inhibitor lactacystin (LC) partially blocked CD4 degradation in transfected HeLa cells coexpressing CD4, Vpu, and HIV-1 Env glycoproteins. In the present study, we show that Vpu-induced CD4 degradation can be completely blocked by proteasome inhibitors, does not require the ER chaperone calnexin, but requires the function of the cytosolic polyubiquitination machinery which apparently targets potential ubiquitination sites within the CD4 cytoplasmic tail. Our findings point to differences between the mechanism of Vpu-mediated CD4 degradation and ERAD processes induced by the HCMV proteins US2 and US11 (61, 62).  相似文献   

5.
6.
Mitotic division requires highly regulated morphological and biochemical changes to the cell. Upon commitment to exit mitosis, cells begin to remove mitotic regulators in a temporally and spatially controlled manner to bring about the changes that reestablish interphase. Ubiquitin-dependent pathways target these regulators to generate polyubiquitin-tagged substrates for degradation by the 26S proteasome. However, the lack of cell-based assays to investigate in vivo ubiquitination limits our knowledge of the identity of substrates of ubiquitin-mediated regulation in mitosis. Here we report an in vivo ubiquitin tagging system used in human cells that allows efficient purification of ubiquitin conjugates from synchronized cell populations. Coupling purification with mass spectrometry, we have identified a series of mitotic regulators targeted for polyubiquitination in mitotic exit. We show that some are new substrates of the anaphase-promoting complex/cyclosome and validate KIFC1 and RacGAP1/Cyk4 as two such targets involved respectively in timely mitotic spindle disassembly and cell spreading. We conclude that in vivo biotin tagging of ubiquitin can provide valuable information about the role of ubiquitin-mediated regulation in processes required for rebuilding interphase cells.Ubiquitination has emerged as a major post-translational modification determining the fate of cellular proteins. One of these fates is proteolysis, whereby the assembly of polyubiquitin chains creates signatures on target proteins that specify delivery to the 26S proteasome for proteolytic destruction. Targeted proteolysis is critical to the control of cell division. For example, the universally conserved mechanism of mitotic exit depends upon rapid proteolysis of mitotic cyclins and securins to drive the transition from mitosis to interphase. This transition is under surveillance by the spindle assembly checkpoint (SAC),1 which controls the activity of a multi-subunit ubiquitin ligase, the anaphase-promoting complex/cyclosome (APC/C) (1, 2).Much of the known specificity in the ubiquitin-proteasome system (UPS) is mediated at the level of substrate targeting by ubiquitin ligase (E3) enzymes, of which there are more than 600 in human cells. Given these facts, it is perhaps surprising that the APC/C is almost the only known engineer of the protein landscape after anaphase onset, targeting mitotic regulators for destruction with high temporal specificity (24). Some roles for nondegradative ubiquitination in regulating the localization of mitotic kinases Aurora B and Plk1 have been described (59), and a growing list of reported ubiquitin interactors can modulate ubiquitin-dependent events during mitosis (10). However, the majority of ubiquitination events that have so far been described as occurring at the transition from mitosis to interphase are APC/C-dependent.Two co-activator subunits, Cdc20 and Cdh1, play vital roles in APC/C-dependent substrate recognition (11) by recognizing two widely characterized degrons, the D-box and the KEN motif (12, 13). Computational approaches that have been used to calculate the total number of APC/C substrates from the prevalence of degrons in the human proteome estimate that there are between 100 and 200 substrates (14), and experiments using in vitro ubiquitination of protein arrays have given rise to estimates in the same range (15). Most of the mitotic regulators targeted by the APC/C during mitotic exit in human cells have been identified via in vitro degradation assays or ubiquitination assays on in vitro–expressed pools of substrates (1518). These approaches have identified several important substrates, but in the absence of in vivo parameters they may not identify substrates whose targeting depends on post-translational modifications or substrates that are only recognized in vivo as components of higher-order complexes. Not all substrates identified in this way have been validated as polyubiquitinated proteins in vivo. Multiple recent proteomic studies have identified large numbers of in vivo ubiquitin-modified sites from yeast (1921) and human cells (2229). None of these studies have used synchronized cell populations to provide information on the timing or regulation of substrate ubiquitination.We reasoned that a better view of ubiquitin-mediated processes that regulate mitotic exit would come from identifying proteins that are ubiquitinated in vivo during mitotic exit. With this goal in mind we adopted a system for in vivo tagging of ubiquitin chains with biotin, previously used to identify ubiquitin-conjugated proteins from the Drosophila neural system (30), and applied it to a human cell line (U2OS) that can be tightly synchronized at mitosis. In contrast to several recent studies that employed antibodies specific to the diGly-Lys remnant that marks ubiquitination sites following trypsin digestion (19, 25), an in vivo ubiquitin tagging strategy allows direct validation of candidate ubiquitinated proteins (whether mono- or polyubiquitinated) through immunoblotting of samples. Moreover, in contrast to other methods for affinity tagging of ubiquitin, or affinity purification via ubiquitin-binding domains, the use of the biotin tag enables purification under highly denaturing conditions for stringent isolation of ubiquitin-conjugated material from higher eukaryotes. His6-tagged ubiquitin is also available for use under denaturing conditions, but it is not generally useful in higher eukaryotic cells, where a high frequency of proteins containing multiple histidine residues confounds the specificity of nickel-affinity pulldowns (as discussed in detail in Ref. 30). Therefore, in this paper we describe the reproducible identification and validation of mitoticphase-specific polyubiquitinated proteins via the in vivo biotinylation of ubiquitin. A large number of polyubiquitinated proteins that we identified are specific to mitotic exit, when the APC/C is active, and we expect that many of them are substrates for the APC/C. We formally identified KIFC1/HSET and Cyk4/RACGAP1 as targets of APC/C-dependent ubiquitin-mediated proteolysis after anaphase onset and investigated the role of their ubiquitination in the regulation of mitotic exit. Cell cycle phase-specific information on protein ubiquitination and the generation of ubiquitinated protein networks provides a framework for further investigation of ubiquitin-controlled processes occurring during the rebuilding of interphase cells.  相似文献   

7.
8.
9.
Posttranslational modifications of proteins increase the complexity of the cellular proteome and enable rapid regulation of protein functions in response to environmental changes. Protein ubiquitylation is a central regulatory posttranslational modification that controls numerous biological processes including proteasomal degradation of proteins, DNA damage repair and innate immune responses. Here we combine high-resolution mass spectrometry with single-step immunoenrichment of di-glycine modified peptides for mapping of endogenous putative ubiquitylation sites in murine tissues. We identify more than 20,000 unique ubiquitylation sites on proteins involved in diverse biological processes. Our data reveals that ubiquitylation regulates core signaling pathways common for each of the studied tissues. In addition, we discover that ubiquitylation regulates tissue-specific signaling networks. Many tissue-specific ubiquitylation sites were obtained from brain highlighting the complexity and unique physiology of this organ. We further demonstrate that different di-glycine-lysine-specific monoclonal antibodies exhibit sequence preferences, and that their complementary use increases the depth of ubiquitylation site analysis, thereby providing a more unbiased view of protein ubiquitylation.Ubiquitin is a small 76-amino-acid protein that is conjugated to the ε-amino group of lysines in a highly orchestrated enzymatic cascade involving ubiquitin activating (E1), ubiquitin conjugating (E2), and ubiquitin ligase (E3) enzymes (1). Ubiquitylation is involved in the regulation of diverse cellular processes including protein degradation (2, 3, 4), DNA damage repair (5, 6), DNA replication (7), cell surface receptor endocytosis, and innate immune signaling (8, 9). Deregulation of protein ubiquitylation is implicated in the development of cancer and neurodegenerative diseases (10, 11). Inhibitors targeting the ubiquitin proteasome system are used in the treatment of hematologic malignancies such as multiple myeloma (12, 13).Recent developments in the mass spectrometry (MS)-based proteomics have greatly expedited proteome-wide analysis of posttranslational modifications (PTMs) (1417). Large-scale mapping of ubiquitylation sites by mass spectrometry is based on the identification of the di-glycine remnant that results from trypsin digestion of ubiquitylated proteins and remains attached to ubiquitylated lysines (18). Recently, two monoclonal antibodies were developed that specifically recognize di-glycine remnant modified peptides enabling their efficient enrichment from complex peptide mixtures (19, 20). These antibodies have been used to identify thousands of endogenous ubiquitylation sites in human cells, and to quantify site-specific changes in ubiquitylation in response to different cellular perturbations (2022). It should be noted that the di-glycine remnant is not specific for proteins modified by ubiquitin but also proteins modified by NEDD8 or ISG15 generate an identical di-glycine remnant on modified lysines making it impossible to distinguish between these modifications by mass spectrometry. However, expression of NEDD8 in mouse tissues was shown to be developmentally down-regulated (23), and ISG15 expression in bovine tissues is low in the absence of interferon stimulation (24). In cell culture experiments it was shown that a great majority of sites identified using di-glycine-lysine-specific antibodies stems from ubiquitylated peptides (20).The rates of cell proliferation and protein turnover in mammals vary dramatically between different tissues. Immortalized cell lines, often derived from cancer, are selected for high proliferation rates and fail to represent the complex conditions in tissues. Tissue proteomics can help to gain a more comprehensive understanding of physiological processes in multicellular organisms. Analysis of tissue proteome and PTMs can provide important insights into tissue-specific processes and signaling networks that regulate these processes (2532). In addition, development of mass spectrometric methods for analysis of PTMs in diseased tissues might lead to the identification of biomarkers.In this study, we combined high-resolution mass spectrometry with immunoenrichment of di-glycine modified peptides to investigate endogenous ubiquitylation sites in murine tissues. We identified more than 20,000 ubiquitylation sites from five different murine tissues and report the largest ubiquitylation dataset obtained from mammalian tissues to date. Furthermore, we compared the performance of the two monoclonal di-glycine-lysine-specific antibodies available for enrichment of ubiquitylated peptides, and reveal their relative preferences for different amino acids flanking ubiquitylation sites.  相似文献   

10.
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   

11.
The Major histocompatibility complex (MHC) class I peptidome is thought to be generated mostly through proteasomal degradation of cellular proteins, a notion that is based on the alterations in presentation of selected peptides following proteasome inhibition. We evaluated the effects of proteasome inhibitors, epoxomicin and bortezomib, on human cultured cancer cells. Because the inhibitors did not reduce the level of presentation of the cell surface human leukocyte antigen (HLA) molecules, we followed their effects on the rates of synthesis of both HLA peptidome and proteome of the cells, using dynamic stable isotope labeling in tissue culture (dynamic-SILAC). The inhibitors reduced the rates of synthesis of most cellular proteins and HLA peptides, yet the synthesis rates of some of the proteins and HLA peptides was not decreased by the inhibitors and of some even increased. Therefore, we concluded that the inhibitors affected the production of the HLA peptidome in a complex manner, including modulation of the synthesis rates of the source proteins of the HLA peptides, in addition to their effect on their degradation. The collected data may suggest that the current reliance on proteasome inhibition may overestimate the centrality of the proteasome in the generation of the MHC peptidome. It is therefore suggested that the relative contribution of the proteasomal and nonproteasomal pathways to the production of the MHC peptidome should be revaluated in accordance with the inhibitors effects on the synthesis rates of the source proteins of the MHC peptides.The repertoires and levels of peptides, presented by the major histocompatibility complex (MHC)1 class I molecules at the cells'' surface, are modulated by multiple factors. These include the rates of synthesis and degradation of their source proteins, the transport efficacy of the peptides through the transporter associated with antigen processing (TAP) into the endoplasmic reticulum (ER), their subsequent processing and loading onto the MHC molecules within the ER, and the rates of transport of the MHC molecules with their peptide cargo to the cell surface. The off-rates of the presented peptides, the residence time of the MHC complexes at the cell surface, and their retrograde transport back into the cytoplasm, influence, as well, the presented peptidomes (reviewed in (1)). Even though significant portions of the MHC class I peptidomes are thought to be derived from newly synthesized proteins, including misfolded proteins, defective ribosome products (DRiPs), and short lived proteins (SLiPs), most of the MHC peptidome is assumed to originate from long-lived proteins, which completed their functional cellular roles or became defective (retirees), (reviewed in (2)).The main protease, supplying the MHC peptidome production pipeline, is thought to be the proteasome (3). It is also responsible for generation of the final carboxyl termini of the MHC peptides (4), (reviewed in (5)). The final trimming of the n-termini of the peptides, within the endoplasmic reticulum (ER), is thought to be performed by amino peptidases, such as ERAP1/ERAAP, which fit the peptides into their binding groove on the MHC molecules (6) (reviewed in (7)). Nonproteasomal proteolytic pathways were also suggested as possible contributors to the MHC peptidome, including proteolysis by the ER resident Signal peptide peptidase (8, 9), the cytoplasmic proteases Insulin degrading enzyme (10), Tripeptidyl peptidase (1113), and a number of proteases within the endolysosome pathway (reviewed recently in (1417)). In contrast to the mostly cytoplasmic and ER production of the MHC class I peptidome, the class II peptidome is produced in a special compartment, associated with the endolysosome pathway (1820). This pathway is also thought to participate in the cross presentation of class I peptides, derived from proteins up-taken by professional antigen presenting cells (21), (reviewed in (1517, 22)).The centrality of the proteasomes in the generation of the MHC peptidome was deduced mostly from the observed change in presentation levels of small numbers of selected peptides, following proteasome inhibition (3, 23). Even the location of some of the genes encoding the catalytic subunits of the immunoproteasome (LMP2 and LMP7) (24) within the MHC class II genomic locus, was suggested to support the involvement of the proteasome in the generation of the MHC class I peptidome (25). Similar conclusions were deduced from alterations in peptide presentation, following expression of the catalytic subunits of the immunoproteasome (26), (reviewed in (5)). Yet, although most of the reports indicated reductions in presentation of selected peptides by proteasome inhibition (3, 2729), others have observed only limited, and sometimes even opposite effects (23, 3032).The matter is further complicated by the indirect effects of proteasome inhibition used for such studies on the arrest of protein synthesis by the cells (3335), on the transport rates of the MHC molecules to the cell surface, and on their retrograde transport back to the vesicular system (36) (reviewed in (37)). Proteasome inhibition likely causes shortage of free ubiquitin, reduced supply of free amino acids, and induces an ER unfolded protein response (UPR), which signals the cells to block most (but not all) cellular protein synthesis (reviewed in (38)). Because a significant portion of the MHC peptidome originates from degradation of DRiPs and SLiPs (reviewed in (2)), arrest of new protein synthesis should influence the presentation of their derived MHC peptides. Taken together, these arguments may suggest that merely following the changes in the presentation levels of the MHC molecules, or even of specific MHC peptides, after proteasome inhibition, does not provide the full picture for deducing the relative contribution of the proteasomal pathway to the production of the MHC peptidome (reviewed in (7)).MHC peptidome analysis can be performed relatively easily by modern capillary chromatography combined with mass spectrometry (reviewed in (39)). The peptides are recovered from immunoaffinity purified MHC molecules after detergent solubilization of the cells (40, 41), from soluble MHC molecules secreted to the cells'' growth medium (42, 43) or from patients'' plasma (44). The purified peptides pools are resolved by capillary chromatography and the individual peptides are identified and quantified by tandem mass spectrometry (40), (reviewed in (4547)). In cultured cells, quantitative analysis can also be followed by metabolic incorporation of stable isotope labeled amino acids (SILAC) (48). Furthermore, the rates of de novo synthesis of both MHC peptides and their proteins of origin can be followed using the dynamic-SILAC proteomics approach (49) with its further adaptation to HLA peptidomics (5052).This study attempts to define the relative contribution of the proteasomes to the production of HLA class I peptidome by simultaneously following the effects of proteasome inhibitors, epoxomicin and bortezomib (Velcade), on the rates of de novo synthesis of both the HLA class I peptidome and the cellular proteome of the same MCF-7 human breast cancer cultured cells. The proteasome inhibitors did not reduce the levels of HLA presentations, yet affected the rates of production of both the HLA peptidome and cellular proteome, mostly decreasing, but also increasing some of the synthesis rates of the HLA peptides and cellular proteins. Thus, we suggest that the degree of contribution of the proteasomal pathway to the production of the HLA-I peptidome should be re-evaluated in accordance with their effects on the entire HLA class-I peptidome, while taking into consideration the inhibitors'' effects on the synthesis (and degradation) rates of the source proteins of each of the studied HLA peptides.  相似文献   

12.
In Archaea, an hexameric ATPase complex termed PAN promotes proteins unfolding and translocation into the 20 S proteasome. PAN is highly homologous to the six ATPases of the eukaryotic 19 S proteasome regulatory complex. Thus, insight into the mechanism of PAN function may reveal a general mode of action mutual to the eukaryotic 19 S proteasome regulatory complex. In this study we generated a three-dimensional model of PAN from tomographic reconstruction of negatively stained particles. Surprisingly, this reconstruction indicated that the hexameric complex assumes a two-ring structure enclosing a large cavity. Assessment of distinct three-dimensional functional states of PAN in the presence of adenosine 5′-O-(thiotriphosphate) and ADP and in the absence of nucleotides outlined a possible mechanism linking nucleotide binding and hydrolysis to substrate recognition, unfolding, and translocation. A novel feature of the ATPase complex revealed in this study is a gate controlling the “exit port” of the regulatory complex and, presumably, translocation into the 20 S proteasome. Based on our structural and biochemical findings, we propose a possible model in which substrate binding and unfolding are linked to structural transitions driven by nucleotide binding and hydrolysis, whereas translocation into the proteasome only depends upon the presence of an unfolded substrate and binding but not hydrolysis of nucleotide.In eukaryotic cells most protein breakdown in the cytosol and nucleus is catalyzed by the 26 S proteasome. This ∼2.5-MDa (1) complex degrades ubiquitin-conjugated and certain non-ubiquitinated proteins in an ATP-dependent manner (2, 3). The 26 S complex is composed of one or two 19 S regulatory particles situated at the ends of the cylindrical 20 S proteasome. Within the 26 S complex, proteins are hydrolyzed in the 20 S proteasome. Tagged substrates, however, first bind to the 19 S regulatory particle, which catalyzes their unfolding and translocation into the 20 S subcomplex (4, 5). The 19 S regulatory particle consists of at least 17 different subunits (1, 6). Nine of these subunits form a “lid,” whereas the other eight subunits, including six ATPases, comprise the base of the 19 S particle. Electron microscopy (710) as well as cross-linking experiments (11, 12) have demonstrated that the six homologous ATPases are associated with the α rings of the 20 S particle.Unlike eukaryotes, Archaea and certain eubacteria contain homologous 20 S particles but lack ubiquitin. Their proteasomes degrade proteins in association with a hexameric ATPase ring complex termed PAN (13). PAN appears to be the evolutionary precursor of the 19 S base, predating the coupling of ubiquitination and proteolysis in eukaryotes (14). In addition, PAN recognizes the bacterial targeting sequence ssrA (in analogy to the polyubiquitin conjugates in eukaryotes) and efficiently unfolds and translocates globular substrates, like green fluorescent protein, when tagged with ssrA (15). In both PAN and the 19 S proteasome regulatory complexes, ATP is essential for substrate unfolding and translocation and for opening of the gated channel in the α ring through which substrates enter the 20 S particle (1517). Because this portal is quite narrow (1820), only extended polypeptides can enter the 20 S proteasome. Consequently, a globular substrate must be unfolded by the associated ATPase complex to be translocated and digested within the 20 S particle.PAN and the six ATPases found at the base of the 19 S particle are members of the AAA+ superfamily of multimeric ATPases which also includes the ATP-dependent proteases Lon and FtsH and the regulatory components of the bacterial ATP-dependent proteases ClpAP, ClpXP, and HslUV (8, 21). For mechanistic studies of the roles of ATP, the simpler archaeal PAN-20 S system offers many technical advantages over the much more complex 26 S proteasome. For example, prior studies of PAN (17, 22) demonstrated that unfolding of globular substrates (e.g. green fluorescent protein-ssrA) requires ATP hydrolysis. The same was also shown for the Escherichia coli ATP-dependent proteases ClpXP (23) and ClpAP (24). We have also shown that unfolding by PAN can take place on the surface of the ATPase ring in the absence of translocation (15). Thus, unfolding seems to proceed independently from protein translocation into the 20 S proteolytic particle. It is noteworthy that other studies suggest that proteins are unfolded by energy-dependent translocation through the ATPase ring (25, 26). These studies have suggested that the translocation of an unfolded polypeptide from the ATPase into the 20 S core is an active process that is coupled to ATP hydrolysis. A key to underline a detailed molecular mechanism for substrate binding, unfolding, and translocation by the proteasome regulatory ATPase complex is improved understanding of its architecture and the nucleotide-dependent structural transitions that afford these functions.To date we and others have failed to generate micrographs suitable for three-dimensional reconstruction of PAN using single-particle EM analysis. Likewise, structural information regarding the three-dimensional architecture and subunit organization within the 19 S particle is very limited. In fact, high resolution three-dimensional information on the 19 S complex is not yet available. Most knowledge available is based on cross-linking experiments (11, 12) as well as EM structural analysis (710), which provided a three-dimensional model outline of the general architecture of the 26 S complex. Unlike the 19 S complex, the structure of the 20 S subcomplex was determined by x-ray crystallography (18, 19). In contrast to the highly homogenous structure of the 20 S complex, the structural heterogeneity and flexibility of the 19 S subcomplex is presumably reflected in multiple conformations, which in turn also contribute to the difficulty in generating a high resolution three-dimensional structural model of the 26 S proteasome. Accordingly, the initial goal of this study was to generate a three-dimensional model of PAN that will allow us to determine its general architecture and to correlate unique conformational transitions within this ATPase with the nucleotide state of the complex (i.e. in the presence of ATPγS, ADP, or in the absence of nucleotides).Smith et al. (27) suggested a general architecture for the PAN-20 S complex based on two-dimensional averaging of a Thermoplasma acidophilum (TA)3 20 S proteasome and Methanococcus jannaschii (MJ) PAN hybrid complex in the presence of ATPγS. Based on side-view projections of that complex, these authors proposed that PAN assumes an overall structure similar to E. coli HslU (2830).We realized that although PAN appears heterogeneous in electron micrographs, it does not occupy all possible orientations when adsorbed to carbon-coated electron microscopy (EM) grids, a prerequisite for single particle analysis. This problem was overcome by applying electron tomography in conjunction with a three-dimensional averaging procedure that accounts for the missing wedge in the Fourier space of electron tomograms (31, 32). The three-dimensional model generated revealed an unexpected architecture leading to a possible molecular mechanism describing the function of PAN and presumably the 19 S ATPases.  相似文献   

13.
Damaged and misfolded proteins that are no longer functional in the cell need to be eliminated. Failure to do so might lead to their accumulation and aggregation, a hallmark of many neurodegenerative diseases. Protein quality control pathways play a major role in the degradation of these proteins, which is mediated mainly by the ubiquitin proteasome system. Despite significant focus on identifying ubiquitin ligases involved in these pathways, along with their substrates, a systems-level understanding of these pathways has been lacking. For instance, as misfolded proteins are rapidly ubiquitylated, unconjugated ubiquitin is rapidly depleted from the cell upon misfolding stress; yet it is unknown whether certain targets compete more efficiently to be ubiquitylated. Using a system-wide approach, we applied statistical and computational methods to identify characteristics enriched among proteins that are further ubiquitylated after heat shock. We discovered that distinct populations of structured and, surprisingly, intrinsically disordered proteins are prone to ubiquitylation. Proteomic analysis revealed that abundant and highly structured proteins constitute the bulk of proteins in the low-solubility fraction after heat shock, but only a portion is ubiquitylated. In contrast, ubiquitylated, intrinsically disordered proteins are enriched in the low-solubility fraction after heat shock. These proteins have a very low abundance in the cell, are rarely encoded by essential genes, and are enriched in binding motifs. In additional experiments, we confirmed that several of the identified intrinsically disordered proteins were ubiquitylated after heat shock and demonstrated for two of them that their disordered regions are important for ubiquitylation after heat shock. We propose that intrinsically disordered regions may be recognized by the protein quality control machinery and thereby facilitate the ubiquitylation of proteins after heat shock.Cells face the constant threat of protein misfolding and aggregation, and thus protein quality control pathways are important in selectively targeting damaged and misfolded proteins for degradation (1, 2). The ubiquitin proteasome system serves as a major mediator of this pathway by conjugating the small protein ubiquitin onto substrates through the E1-E2-E3 (ubiquitin-activating enzyme, ubiquitin-conjugating enzyme, and ubiquitin ligase, respectively) cascade for their recognition and degradation by the proteasome (3, 4). It is known that the activity of the ubiquitin-proteasome system is associated with many neurodegenerative diseases. For instance, ubiquitin is found enriched in protein inclusions associated with these diseases (5). Furthermore, proteasome activity has been shown to decrease with age in a large variety of organisms (6), leading to increased proteotoxicity in the cell.Because of the importance of maintaining protein homeostasis, numerous ubiquitin ligases in different cellular compartments function in protein quality control pathways to target misfolded or damaged proteins for degradation via the proteasome. For instance, the conserved Hrd1 ubiquitin ligase is involved in the endoplasmic-reticulum-associated degradation pathway that targets endoplasmic reticulum proteins for retro-translocation to the cytoplasm and proteasome degradation (7). A major question is what features are recognized by ubiquitin ligases that allow them to selectively target terminally misfolded proteins for degradation, given that the folding rates and physicochemical properties vary largely from protein to protein. Several E3 ubiquitin ligases involved in cytosolic protein quality control target their substrates via their interactions with chaperone proteins. For instance, the CHIP ubiquitin ligase can directly bind to Hsp70 and Hsp90 proteins (8), which may hand over client proteins that are not successfully folded. Understanding which features are recognized by these degradation quality-control pathways might help us understand how certain misfolded proteins evade this system, leading to their accumulation and aggregation in the cell.Many studies investigating degradation protein quality control have employed model substrates (e.g. mutated proteins that misfold) to reveal which components are involved in a given quality control machinery. However, these approaches do not typically reveal the whole spectrum of substrates for these pathways. Thus, alternative system-wide approaches are also needed to provide a bigger picture. Heat shock (HS)1 induces general misfolding at the proteome level by increasing thermal energy and was shown to cause an increase in ubiquitylation levels in the cell over 25 years ago (9, 10). However, the exact mechanism and pathways that target misfolded proteins have remained uncharacterized for a long time. We recently showed that the Hul5 ubiquitin ligase plays a major role in this heat stress response that mainly affects cytosolic proteins (11). Absence of Hul5 averts the ubiquitylation in the cytoplasm of several misfolded targets after HS, as well as low-solubility proteins in unstressed cells. Other E3 ubiquitin ligases are likely involved in this pathway (12). Interestingly, as ubiquitin constitutes about only 1% of the proteome, free unconjugated ubiquitin is rapidly depleted under stress conditions (13, 14). Given the limited amount of this protein, how does the cell triage ubiquitin among an excess of misfolded proteins? In order to gain systems-level insight, we sought to identify characteristics enriched among proteins ubiquitylated after HS using a combination of statistical and computational analysis, and we conducted additional proteomics and biochemical experiments to support our hypotheses. We discovered an unexpected susceptibility of intrinsically disordered proteins for ubiquitylation after misfolding stress.  相似文献   

14.
Protein degradation provides an important regulatory mechanism used to control cell cycle progression and many other cellular pathways. To comprehensively analyze the spatial control of protein degradation in U2OS osteosarcoma cells, we have combined drug treatment and SILAC-based quantitative mass spectrometry with subcellular and protein fractionation. The resulting data set analyzed more than 74,000 peptides, corresponding to ∼5000 proteins, from nuclear, cytosolic, membrane, and cytoskeletal compartments. These data identified rapidly degraded proteasome targets, such as PRR11 and highlighted a feedback mechanism resulting in translation inhibition, induced by blocking the proteasome. We show this is mediated by activation of the unfolded protein response. We observed compartment-specific differences in protein degradation, including proteins that would not have been characterized as rapidly degraded through analysis of whole cell lysates. Bioinformatic analysis of the entire data set is presented in the Encyclopedia of Proteome Dynamics, a web-based resource, with proteins annotated for stability and subcellular distribution.Targeted protein degradation is an important regulatory mechanism that allows co-ordination of cellular pathways in response to environmental and temporal stimuli (1). The control of diverse biochemical pathways, including cell cycle progression and the response to DNA damage, is mediated, at least in part, by dynamic alterations in protein degradation (2). Previous large scale proteomics studies in mammalian cells have shown that the rate of protein degradation can vary from the timescale of minutes, to essentially infinite stability for metastable proteins (38).Most intracellular proteins have similar degradation rates, with a half-life approximating the cell doubling rate. Under 5% of proteins display degradation rates more than threefold faster than the proteome average (35, 7). However, degradation rates for individual proteins can change, for example depending on either the cell cycle stage, or signaling events, and can also vary depending on subcellular localization. Disruption of such regulated protein stability underlies the disease mechanisms responsible for forms of cancer, e.g. p53 (9, 10) and the proto-oncogene c-Myc (11).Detection of rapidly degraded proteins can be difficult because of their low abundance. However, advances in mass spectrometry based proteomics have enabled in-depth quantitative analysis of cellular proteomes (1214). Stable isotope labeling by amino acids in cell culture (SILAC)1 (15), has been widely used to measure protein properties such as abundance, interactions, modifications, turnover, and subcellular localization under different conditions (16). Subcellular fractionation and protein size separation are also powerful techniques that enhance in-depth analysis of cellular proteomes. Not only do these fractionation techniques increase total proteome coverage, they also provide biological insight regarding how protein behavior differs between subcellular compartments. For example, subcellular fractionation has highlighted differences in the rate of ribosomal protein degradation between the nucleus and cytoplasm, (7, 17). Other studies have also demonstrated the benefit of in-depth subcellular fractionation and created methods for the characterization of how proteomes are localized in organelles (1820).In this study we have used SILAC-based quantitative mass spectrometry combined with extensive subcellular and protein-level fractionation to identify rapidly degraded proteins in human U2OS cells. We provide a proteome level characterization of a major feedback mechanism involving inhibition of protein translation when the proteasome is inhibited. We also present the Encyclopedia of Proteome Dynamics, a user-friendly online resource providing access to the entire data set.  相似文献   

15.
16.
17.
ATP-dependent proteases control the concentrations of hundreds of regulatory proteins and remove damaged or misfolded proteins from cells. They select their substrates primarily by recognizing sequence motifs or covalent modifications. Once a substrate is bound to the protease, it has to be unfolded and translocated into the proteolytic chamber to be degraded. Some proteases appear to be promiscuous, degrading substrates with poorly defined targeting signals, which suggests that selectivity may be controlled at additional levels. Here we compare the abilities of representatives from all classes of ATP-dependent proteases to unfold a model substrate protein and find that the unfolding abilities range over more than 2 orders of magnitude. We propose that these differences in unfolding abilities contribute to the fates of substrate proteins and may act as a further layer of selectivity during protein destruction.Energy-dependent proteolysis is responsible for more than 90% of the protein turnover inside the cell (1). This process both removes misfolded and aggregated proteins as part of the response of the cell to stress and controls the concentrations of regulatory proteins (2, 3). In prokaryotes and eukaryotic organelles, energy-dependent proteases fall into five classes as follows: ClpAP, ClpXP, Lon, HslUV (also referred to as ClpYQ), and HflB (also referred to as FtsH). In Archaea, analogous functions are performed by the archaebacterial proteasome, consisting of the proteasome-activating nucleotidase (PAN),3 working with the 20 S proteasome (4); in the cytoplasm and nucleus of eukaryotes, these same functions are performed by the 26 S proteasome (5). These different proteases show little sequence conservation outside the ATP-binding domains, but they share their overall architecture. They all form oligomeric, barrel-shaped complexes composed of one or more rings with the active sites of proteolysis sequestered inside a central degradation chamber (6). Access channels to these sites are narrow, and proteins have to be unfolded to gain entry (6). Regulatory particles belonging to the AAA family of molecular chaperones assemble on either end of the proteolytic chamber and recognize substrates destined for degradation. After recognition, the regulatory particles translocate the substrate through a central channel to the proteolytic chamber and in doing so unravel folded domains within the substrate. Translocation and unfolding are driven by ATP hydrolysis by the regulatory particles, with conformational changes in the protease transmitted to the substrate by conserved residues in the loops lining the channel (710).Protein degradation by AAA proteases is tightly regulated. Most proteins are targeted to ClpAP, ClpXP, HslUV, Lon, HflB, and PAN by sequence motifs in their primary structure (1117). Sometimes adaptor proteins recognize and bind sequence elements in substrates and deliver them to the protease, and other times the protease recognizes sequence elements directly (18, 19). In contrast, proteins are typically targeted to the 26 S proteasome through the covalent attachment of polyubiquitin chains (20). Thus, substrates appear to be selected for degradation based on the presence of specific recognition elements in the protein substrates.However, other mechanisms may also affect the specificity of degradation by prokaryotic proteases. Individual proteases recognize a wide range of targeting signals (11, 16). (For example, Escherichia coli ClpXP recognizes sequences belonging to five distinct classes of consensus sequences (11), and ClpAP, Lon, and FtsH can bind to unstructured regions in proteins with a wide range of amino acid sequences (2123).) One illustration of the loose specificity in targeting signals is the ability of a mitochondrial presequence to target proteins to the proteases ClpAP (24) and HslUV in vitro (see below). In addition, substrates are commonly acted upon by several different proteases in E. coli. For instance, proteins containing the 11-residue ssrA peptide at their C termini can be recognized by ClpAP, ClpXP, FtsH, Lon, and the archaebacterial proteasome (4, 2527). Similarly, some substrates of Lon can be degraded by HslUV in vivo (28).It is not clear how degradation remains selective despite the loose specificity of targeting signals. We propose that the intrinsic protein unfolding ability of AAA proteases and the stabilities of substrates against unfolding play a role in determining the fate of cellular proteins. For example, ClpXP releases hard-to-unfold substrates when it encounters them and degrades destabilized titin variants 20-fold faster than wild type titin (29). The membrane-bound AAA protease FtsH has a weak unfolding ability, which allows this protease to act selectively on damaged and unfolded polypeptides (30). Here we find that the relative unfolding abilities of ATP-dependent proteases vary more than 100-fold and that the unfolding abilities of proteases belonging to the same class but originating from different species appear to be conserved. The unfolding abilities also seem to be intrinsic properties of the proteases themselves rather than other cytosolic factors, such as chaperones. Differences in protease unfolding abilities may contribute to substrate selectivity during protein degradation. For example, expression of a protease with a weak unfolding ability during a stress response could allow the selective elimination of unfolded, misfolded, or otherwise aberrant proteins and spare stable proteins from destruction (30).  相似文献   

18.
19.
20.
The cystic fibrosis transmembrane conductance regulator (CFTR), a member of the ABC transporter superfamily, is a cyclic AMP-regulated chloride channel and a regulator of other ion channels and transporters. In epithelial cells CFTR is rapidly endocytosed from the apical plasma membrane and efficiently recycles back to the plasma membrane. Because ubiquitination targets endocytosed CFTR for degradation in the lysosome, deubiquitinating enzymes (DUBs) are likely to facilitate CFTR recycling. Accordingly, the aim of this study was to identify DUBs that regulate the post-endocytic sorting of CFTR. Using an activity-based chemical screen to identify active DUBs in human airway epithelial cells, we demonstrated that Ubiquitin Specific Protease-10 (USP10) is located in early endosomes and regulates the deubiquitination of CFTR and its trafficking in the post-endocytic compartment. small interference RNA-mediated knockdown of USP10 increased the amount of ubiquitinated CFTR and its degradation in lysosomes, and reduced both apical membrane CFTR and CFTR-mediated chloride secretion. Moreover, a dominant negative USP10 (USP10-C424A) increased the amount of ubiquitinated CFTR and its degradation, whereas overexpression of wt-USP10 decreased the amount of ubiquitinated CFTR and increased the abundance of CFTR. These studies demonstrate a novel function for USP10 in facilitating the deubiquitination of CFTR in early endosomes and thereby enhancing the endocytic recycling of CFTR.The endocytosis, endocytic recycling, and endosomal sorting of numerous transport proteins and receptors are regulated by ubiquitination (16). Ubiquitin, an 8-kDa protein, is conjugated to target proteins via a series of steps that includes ubiquitin-activating enzymes (E1),2 ubiquitin-conjugating enzymes (E2), and ubiquitin ligases (E3) (1). Proteins that are ubiquitinated in the plasma membrane are internalized and are either deubiquitinated and recycle back to the plasma membrane or, via interactions with the endosomal sorting complexes required for transport machinery, are delivered to the lysosome for degradation (17). Sorting of ubiquitinated plasma membrane proteins for either the lysosomal pathway or for the recycling pathway is regulated, in part, by the removal of ubiquitin by deubiquitinating enzymes (DUBs) (16). Thus, the balance between ubiquitination and deubiquitination regulates the plasma membrane abundance of several membrane proteins, including the epithelial sodium channel (ENaC), the epidermal growth factor receptor, the transforming growth factor-β receptor, and the cytokine receptor γ-c (814).CFTR is rapidly endocytosed from the plasma membrane and undergoes rapid and efficient recycling back to the plasma membrane in human airway epithelial cells, with >75% of endocytosed wild-type CFTR recycling back to the plasma membrane (1518). A study published several years ago demonstrated that, although ubiquitination did not regulate CFTR endocytosis, ubiquitination reduced the plasma membrane abundance of CFTR in BHK cells by redirecting CFTR from recycling endosomes to lysosomes for degradation (19). However, neither the E3 ubiquitin ligase(s) responsible for the ubiquitination of CFTR nor the DUB(s) responsible for the deubiquitination of CFTR in the endocytic pathway have been identified in any cell type. Moreover, the effect of the ubiquitin status of CFTR on its endocytic sorting in human airway epithelial cells has not been reported. Thus, the goals of this study were to determine if the ubiquitin status regulates the post-endocytic sorting of CFTR in polarized airway epithelial cells, and to identify the DUBs that deubiquitinate CFTR.Approximately 100 DUBs have been identified in the human genome and are classified into five families based on sequence similarity and mechanism of action (16, 20, 21). To identify DUBs that regulate the deubiquitination of CFTR from this large class of enzymes, we chose an activity-based, chemical probe screening approach developed by Dr. Hidde Ploegh (4, 21, 22). This approach utilizes a hemagglutinin (HA)-tagged ubiquitin probe engineered with a C-terminal modification incorporating a thiol-reactive group that forms an irreversible, covalent bond with active DUBs. Using this approach we demonstrated in polarized human airway epithelial cells that ubiquitin-specific protease-10 (USP10) is located in early endosomes and regulates the deubiquitination of CFTR and thus its trafficking in the post-endocytic compartment. These studies demonstrate a novel function for USP10 in promoting the deubiquitination of CFTR in early endosomes and thereby enhancing the endocytic recycling of CFTR.  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