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 93 毫秒
1.
Tautomycin (TTM) is a highly potent and specific protein phosphatase inhibitor isolated from Streptomyces spiroverticillatus. The biological activity of TTM makes it an important lead for drug discovery, whereas its spiroketal-containing polyketide chain and rare dialkylmaleic anhydride moiety draw attention to novel biosynthetic chemistries responsible for its production. To elucidate the biosynthetic machinery associated with these novel molecular features, the ttm biosynthetic gene cluster from S. spiroverticillatus was isolated and characterized, and its involvement in TTM biosynthesis was confirmed by gene inactivation and complementation experiments. The ttm cluster was localized to a 86-kb DNA region, consisting of 20 open reading frames that encode three modular type I polyketide synthases (TtmHIJ), one type II thioesterase (TtmT), five proteins for methoxymalonyl-S-acyl carrier protein biosynthesis (Ttm-ABCDE), eight proteins for dialkylmaleic anhydride biosynthesis and regulation (TtmKLMNOPRS), as well as two additional regulatory proteins (TtmF and TtmQ) and one tailoring enzyme (TtmG). A model for TTM biosynthesis is proposed based on functional assignments from sequence analysis, which agrees well with previous feeding experiments, and has been further supported by in vivo gene inactivation experiments. These findings set the stage to fully investigate TTM biosynthesis and to biosynthetically engineer new TTM analogs.Tautomycin (TTM)2 is a polyketide natural product first isolated in 1987 from Streptomyces spiroverticillatus (1). The structure and stereochemistry of TTM were established on the basis of chemical degradation and spectroscopic evidence (2-4). TTM contains several features not common to polyketide natural products, including a spiroketal group, a methoxymalonate-derived unit, and an acyl chain bearing a dialkylmaleic anhydride moiety. Structurally related to TTM is tautomycetin (TTN), which was first isolated in 1989 from Streptomyces griseochromogenes following the discovery of TTM (5, 6). The structure of TTN was deduced by chemical degradation and spectroscopic analysis (6), and its stereochemistry was established by comparison of spectral data with those of TTN degradation products and synthetic fragments (7). Both TTM and TTN exist as tautomeric mixtures composed of two interconverting anhydride and diacid forms in approximately a 5:4 ratio under neutral conditions (Fig. 1A) (1, 2).Open in a separate windowFIGURE 1.A, structures of TTM and TTN in anhydride or diacid forms, and biosynthetic origin of the dialkylmaleic anhydride by feeding experiments using 13C-labeled acetate and propionate. The methoxymalonate-derived unit in TTM is highlighted by the dotted oval. R, polyketide moiety of TTM or TTN. B, selected natural product inhibitors of PP-1 and PP-2A featuring a spiroketal or dialkylmaleric anhydride moiety. C, selected natural products containing a dialkylmaleic anhydride moiety.Early studies of TTM revealed its ability to induce morphological changes in leukemia cells (8). However, it was later realized that TTM is a potent and specific inhibitor of protein phosphatases (PPs) PP-1 and PP-2A (9). PP-1 and PP-2A are two of the four major serine/threonine protein phosphatases that regulate diverse cellular events such as cell division, gene expression, muscle contraction, glycogen metabolism, and neuronal signaling in eukaryotic cells (10-12). Many natural product PP-1 and PP-2A inhibitors are known, including okadaic acid (13), calyculin-A (14), phoslactomycin, spirastrellolide, and cantharidin (15) (Fig. 1B), as well as TTM (16, 17), and TTN (18). They have served as useful tools to study PP-involved intracellular events in vivo and as novel leads for drug discovery (10-12). Among these PP inhibitors, TTM and TTN are unique because of their PP-1 selectivity. Despite their structural similarities, TTM exhibits potent specific inhibition of PP-1 and PP-2A with IC50 values of 22-32 nm and only a slight preference for PP-1 (18). Conversely, TTN shows nearly a 40-fold higher binding affinity to PP-1 (IC50 = 1.6 nm) than to PP-2A (IC50 = 62 nm) (18). Because the major structural differences between TTM and TTN reside in the region distal to the dialkylmaleic anhydride moiety (Fig. 1A), it has been proposed that differences in these moieties might be responsible for the PP-1 selectivity (17-19). Finally, TTN also has an impressive immuno-suppressive activity (20, 21), which is apparently devoid for TTM. Clearly, the structural differences between these two polyketides translate into large, exploitable differences in bio-activities, yet an understanding of the biosynthetic origins of these differences remains elusive.The spiroketal and dialkylmaleic anhydride features of TTM are uncommon for polyketide natural products, as is the methoxymalonate-derived unit (Fig. 1A). Few studies have been carried out for spiroketal biosynthesis, yet it is reasonably common among the phosphatase inhibitors such as calyculin A, okadaic acid, and a few others (Fig. 1B). Less common, but still found in the phosphatase inhibitor cantharidin, as well as TTM and TTN, is the dialkylmaleic anhydride moiety (Fig. 1B); this unit appears in a number of other natural products (Fig. 1C), although the biosynthetic steps leading to this reactive moiety (a protected version of a dicarboxylate) have not been rigorously investigated. Feeding experiments with 13C-labeled precursors indicated that the anhydride of TTM and TTN is assembled from a propionate and an as yet undefined C-5 unit (Fig. 1A), which would require novel chemistry for polyketide biosynthesis (22). TTM differentiates itself from all known PP-1 and PP-2A inhibitors by virtue of its unique combination of both the dialkymaleic anhydride and spiroketal functionalities.Multiple total syntheses of TTM and a small number of analogs have been reported, confirming the predicted structure and absolute stereochemistry and facilitating structure-activity relationship studies on PP inhibition and apoptosis induction (19, 23-25). These studies revealed that: (i) the C22-C26 carbon chain and the dialkylmaleic anhydride are the minimum requirements for TTM bioactivity; (ii) the C18-C21 carbon chain and 22-hydroxy group are important for PP inhibition; (iii) the spiroketal moiety determines the affinity to specific protein phosphatases; (iv) the active form is most likely the dicarboxylate; and (v) 3′-epi-TTM exhibits 1,000-fold less activity than TTM. However, taken as a whole, none of the analogs had an improved potency or selectivity for PP-1 inhibition than the natural TTM (19, 22-25). As a result, a more specific inhibitor of PP-1 is urgently awaited to differentiate the physiological roles of PP-1 and PP-2A in vivo and to explore PPs as therapeutic targets for drug discovery.We have undertaken the cloning and characterization of the TTM biosynthetic gene cluster from S. spiroverticillatus as the first step toward engineering TTM biosynthesis for novel analogs (26). We report here: (i) cloning and sequencing of the complete ttm gene cluster, (ii) determination of the ttm gene cluster boundaries, (iii) bioinformatics analysis of the ttm cluster and a proposal for TTM biosynthesis, and (iv) genetic characterization of the TTM pathway to support the proposed pathway. Of particular interest has been the identification of genes possibly related to dialkylmaleic anhydride biosynthesis, the unveiling of the ttm polyketide synthase (PKS) genes predicted to select and incorporate four different starter and extender units for TTM production, and the apparent lack of candidate genes associated with spiroketal formation. These findings now set the stage to engineer TTM analogs for novel PP-1- and PP-2A-specific inhibitors by applying combinatorial biosynthetic methods to the TTM biosynthetic machinery.  相似文献   

2.
Archaeal methane formation from methylamines is initiated by distinct methyltransferases with specificity for monomethylamine, dimethylamine, or trimethylamine. Each methylamine methyltransferase methylates a cognate corrinoid protein, which is subsequently demethylated by a second methyltransferase to form methyl-coenzyme M, the direct methane precursor. Methylation of the corrinoid protein requires reduction of the central cobalt to the highly reducing and nucleophilic Co(I) state. RamA, a 60-kDa monomeric iron-sulfur protein, was isolated from Methanosarcina barkeri and is required for in vitro ATP-dependent reductive activation of methylamine:CoM methyl transfer from all three methylamines. In the absence of the methyltransferases, highly purified RamA was shown to mediate the ATP-dependent reductive activation of Co(II) corrinoid to the Co(I) state for the monomethylamine corrinoid protein, MtmC. The ramA gene is located near a cluster of genes required for monomethylamine methyltransferase activity, including MtbA, the methylamine-specific CoM methylase and the pyl operon required for co-translational insertion of pyrrolysine into the active site of methylamine methyltransferases. RamA possesses a C-terminal ferredoxin-like domain capable of binding two tetranuclear iron-sulfur proteins. Mutliple ramA homologs were identified in genomes of methanogenic Archaea, often encoded near methyltrophic methyltransferase genes. RamA homologs are also encoded in a diverse selection of bacterial genomes, often located near genes for corrinoid-dependent methyltransferases. These results suggest that RamA mediates reductive activation of corrinoid proteins and that it is the first functional archetype of COG3894, a family of redox proteins of unknown function.Most methanogenic Archaea are capable of producing methane only from carbon dioxide. The Methanosarcinaceae are a notable exception as representatives are capable of methylotrophic methanogenesis from methylated amines, methylated thiols, or methanol. Methanogenesis from these substrates requires methylation of 2-mercaptoethanesulfonic acid (coenzyme M or CoM) that is subsequently used by methylreductase to generate methane and a mixed disulfide whose reduction leads to energy conservation (14).Methylation of CoM with trimethylamine (TMA),4 dimethylamine (DMA), or monomethylamine (MMA) is initiated by three distinct methyltransferases that methylate cognate corrinoid-binding proteins (3). MtmB, the MMA methyltransferase, specifically methylates cognate corrinoid protein, MtmC, with MMA (see Fig. 1) (5, 6). The DMA methyltransferase, MtbB, and its cognate corrinoid protein, MtbC, interact specifically to demethylate DMA (7, 8). TMA is demethylated by the TMA methyltransferase (MttB) in conjunction with the TMA corrinoid protein (MttC) (8, 9). Each of the methylated corrinoid proteins is a substrate for a methylcobamide:CoM methyltransferase, MtbA, which produces methyl-CoM (1012).Open in a separate windowFIGURE 1.MMA:CoM methyl transfer. A schematic of the reactions catalyzed by MtmB, MtmC, and MtbA is shown that emphasizes the key role of MtmC in the catalytic cycle of both methyltransferases. Oxidation to Co(II)-MtmC of the supernucleophilic Co(I)-MtmC catalytic intermediate inactivates methyl transfer from MMA to the thiolate of coenzyme M (HSCoM). In vitro reduction of the Co(II)-MtmC with either methyl viologen reduced to the neutral species or with RamA in an ATP-dependent reaction can regenerate the Co(I) species. In either case in vitro Ti(III)-citrate is the ultimate source of reducing power.CoM methylation with methanol requires the methyltransferase MtaB and the corrinoid protein MtaC, which is then demethylated by another methylcobamide:CoM methyltransferase, MtaA (1315). The methylation of CoM with methylated thiols such as dimethyl sulfide in Methanosarcina barkeri is catalyzed by a corrinoid protein that is methylated by dimethyl sulfide and demethylated by CoM, but in this case an associated CoM methylase carries out both methylation reactions (16).In bacteria, analogous methyltransferase systems relying on small corrinoid proteins are used to achieve methylation of tetrahydrofolate. In Methylobacterium spp., CmuA, a single methyltransferase with a corrinoid binding domain, along with a separate pterin methylase, effect the methylation of tetrahydrofolate with chloromethane (17, 18). In Acetobacterium dehalogenans and Moorella thermoacetica various three-component systems exist for specific demethylation of different phenylmethyl ethers, such as vanillate (19) and veratrol (20), again for the methylation of tetrahydrofolate. Sequencing of the genes encoding the corrinoid proteins central to the archaeal and bacterial methylotrophic pathways revealed they are close homologs. Furthermore, genes predicted to encode such corrinoid proteins and pterin methyltransferases are widespread in bacterial genomes, often without demonstrated metabolic function. All of these corrinoid proteins are similar to the well characterized cobalamin binding domain of methionine synthase (21, 22).In contrast, the TMA, DMA, MMA, and methanol methyltransferases are not homologous proteins. The methylamine methyltransferases do share the common distinction of having in-frame amber codons (6, 8) within their encoding genes that corresponds to the genetically encoded amino acid pyrrolysine (2325). Pyrrolysine has been proposed to act in presenting a methylammonium adduct to the central cobalt ion of the corrinoid protein for methyl transfer (3, 23, 26). However, nucleophilic attack on a methyl donor requires the central cobalt ion of a corrinoid cofactor is in the nucleophilic Co(I) state rather than the inactive Co(II) state (27). Subsequent demethylation of the methyl-Co(III) corrinoid cofactor regenerates the nucleophilic Co(I) cofactor. The Co(I)/Co(II) in the cobalamin binding domain of methionine synthase has an Em value of -525 mV at pH 7.5 (28). It is likely to be similarly low in the homologous methyltrophic corrinoid proteins. These low redox potentials make the corrinoid cofactor subject to adventitious oxidation to the inactive Co(II) state (Fig. 1).During isolation, these corrinoid proteins are usually recovered in a mixture of Co(II) or hydroxy-Co(III) states. For in vitro studies, chemical reduction can maintain the corrinoid protein in the active Co(I) form. The methanol:CoM or the phenylmethyl ether:tetrahydrofolate methyltransferase systems can be activated in vitro by the addition of Ti(III) alone as an artificial reductant (14, 19). In contrast, activation of the methylamine corrinoid proteins further requires the addition of methyl viologen as a redox mediator. Ti(III) reduces methyl viologen to the extremely low potential neutral species. In vitro activation with these agents does not require ATP (5, 7, 9).Cellular mechanisms also exist to achieve the reductive activation of corrinoid cofactors in methyltransferase systems. Activation of human methionine synthase involves reduction of the co(II)balamin by methionine synthase reductase (29), whereas the Escherichia coli enzyme requires flavodoxin (30). The endergonic reduction is coupled with the exergonic methylation of the corrinoid with S-adenosylmethionine (27). An activation system exists in cellular extracts of A. dehalogenans that can activate the veratrol:tetrahydrofolate three-component system and catalyze the direct reduction of the veratrol-specific corrinoid protein to the Co(I) state; however, the activating protein has not been purified (31).For the methanogen methylamine and methanol methyltransferase systems, an activation process is readily detectable in cell extracts that is ATP- and hydrogen-dependent (32, 33). Daas et al. (34, 35) examined the activation of the methanol methyltransferase system in M. barkeri and purified in low yield a methyltransferase activation protein (MAP) which in the presence of a preparation of hydrogenase and uncharacterized proteins was required for ATP-dependent reductive activation of methanol:CoM methyl transfer. MAP was found to be a heterodimeric protein without a UV-visible detectable prosthetic group. Unfortunately, no protein sequence has been reported for MAP, leaving the identity of the gene in question. The same MAP protein was also suggested to activate methylamine:CoM methyl transfer, but this suggestion was based on results with crude protein fractions containing many cellular proteins other than MAP (36).Here we report of the identification and purification to near-homogeneity of RamA (reductive activation of methyltransfer, amines), a protein mediating activation of methylamine:CoM methyl transfer in a highly purified system (Fig. 1). Quite unlike MAP, which was reported to lack prosthetic groups, RamA is an iron-sulfur protein that can catalyze reduction of a corrinoid protein such as MtmC to the Co(I) state in an ATP-dependent reaction (Fig. 1). Peptide mapping of RamA allowed identification of the gene encoding RamA and its homologs in the genomes of Methanosarcina spp. RamA belongs to COG3894, a group of uncharacterized metal-binding proteins found in a number of genomes. RamA, thus, provides a functional example for a family of proteins widespread among bacteria and Archaea whose physiological role had been largely unknown.  相似文献   

3.
In mammals, xylose is found as the first sugar residue of the tetrasaccharide GlcAβ1-3Galβ1-3Galβ1-4Xylβ1-O-Ser, initiating the formation of the glycosaminoglycans heparin/heparan sulfate and chondroitin/dermatan sulfate. It is also found in the trisaccharide Xylα1-3Xylα1-3Glcβ1-O-Ser on epidermal growth factor repeats of proteins, such as Notch. UDP-xylose synthase (UXS), which catalyzes the formation of the UDP-xylose substrate for the different xylosyltransferases through decarboxylation of UDP-glucuronic acid, resides in the endoplasmic reticulum and/or Golgi lumen. Since xylosylation takes place in these organelles, no obvious requirement exists for membrane transport of UDP-xylose. However, UDP-xylose transport across isolated Golgi membranes has been documented, and we recently succeeded with the cloning of a human UDP-xylose transporter (SLC25B4). Here we provide new evidence for a functional role of UDP-xylose transport by characterization of a new Chinese hamster ovary cell mutant, designated pgsI-208, that lacks UXS activity. The mutant fails to initiate glycosaminoglycan synthesis and is not capable of xylosylating Notch. Complementation was achieved by expression of a cytoplasmic variant of UXS, which proves the existence of a functional Golgi UDP-xylose transporter. A ∼200 fold increase of UDP-glucuronic acid occurred in pgsI-208 cells, demonstrating a lack of UDP-xylose-mediated control of the cytoplasmically localized UDP-glucose dehydrogenase in the mutant. The data presented in this study suggest the bidirectional transport of UDP-xylose across endoplasmic reticulum/Golgi membranes and its role in controlling homeostasis of UDP-glucuronic acid and UDP-xylose production.Xylose is only known to occur in two different mammalian glycans. First, xylose is the starting sugar residue of the common tetrasaccharide, GlcAβ1,3Galβ1,3Galβ1,4Xylβ1-O-Ser, attached to proteoglycan core proteins to initiate the biosynthesis of glycosaminoglycans (GAGs)2 (1). Second, xylose is found in the trisaccharide Xylα1,3Xylα1,3Glcβ1-O-Ser in epidermal growth factor (EGF)-like repeats of proteins, such as blood coagulation factors VII and IX (2) and Notch (3) (Fig. 1). Two variants of O-xylosyltransferases (XylT1 and XylT2) are responsible for the initiation of glycosaminoglycan biosynthesis, which differ in terms of acceptor specificity and tissue distribution (4-7), and two different enzymatic activities have been identified that catalyze xylosylation of O-glucose residues added to EGF repeats (8-10). On Notch, O-glucose occurs on EGF repeats in a similar fashion as O-fucose, which modifications have been shown to influence ligand-mediated Notch signaling (11-16). Recently, rumi, the gene encoding the Notch O-glucosyltransferase in Drosophila, has been identified, and inactivation of the gene was found to cause a temperature-sensitive Notch phenotype (17). Although this finding clearly demonstrated that O-glucosylation is essential for Notch signaling, the importance of xylosylation for Notch functions remains ambiguous.Open in a separate windowFIGURE 1.UDP-xylose metabolism in mammalian cells. A, UDP-Xyl is synthesized in two steps from UDP-Glc by the enzymes UGDH, forming UDP-GlcA, and UXS, also referred to as UDP-glucuronic acid decarboxylase. UGDH is inhibited by the product of the second enzyme, UDP-Xyl (42). B, in mammals, UDP-Xyl is synthesized within the lumen of the ER/Golgi, where it is substrate for different xylosyltransferases incorporating xylose in the glycosaminoglycan core (XylT1 and XylT2) or in O-glucose-linked glycans. The nucleotide sugar transporter SLC35D1 (52) has been shown to transport UDP-GlcA over the ER membrane and SLC35B4 (29) to transport UDP-Xyl over the Golgi membrane. The function of this latter transporter is unclear.Several different Chinese hamster ovary (CHO) cell lines with defects in GAG biosynthesis have been isolated by screening for reduced incorporation of sulfate (18) and reduced binding of fibroblast growth factor 2 (FGF-2) (19, 20) and by direct selection with FGF-2 conjugated to the plant cytotoxin saporin (21). Isolated cells (called pgs, for proteoglycan synthesis mutants) (21) exhibited defects in various stages of GAG biosynthesis, ranging from the initiating xylosyltransferase to specific sulfation reactions (18, 19, 21-25). Mutants that affect overall GAG biosynthesis were shown to have a defect in the assembly of the common core tetrasaccharide. Interestingly, these latter mutants could be separated into clones in which GAG biosynthesis can be restored by the external addition of xylosides as artificial primers and those that cannot (18). The two mutants belonging to the first group are pgsA-745 and pgsB-761. Although pgs-745 is defective in XylT2 (4-6, 18), pgsB-761 exhibits a defect in galactosyltransferase I (B4GalT7), the enzyme that catalyzes the first step in the elongation of the xylosylated protein (25 (see Fig. 1B). Restoration of GAG biosynthesis in the latter mutant presumably occurs through a second β1-4-galactosyltransferase, able to act on xylosides when provided at high concentration but not on the endogenous protein-linked xylose.Here we describe the isolation of a third CHO cell line (pgsI-208) with the xyloside-correctable phenotype. The mutant is deficient in UDP-xylose synthase (UXS), also known as UDP-glucuronic acid decarboxylase. This enzyme catalyzes the synthesis of UDP-Xyl, the common donor substrate for the different xylosyltransferases, by decarboxylation of UDP-glucuronic acid. Importantly, UXS in the animal cell is localized in the lumen of the ER and/or Golgi (26-28), superseding at first sight the need for the Golgi UDP-xylose transporter, which has been recently cloned and characterized (29). Using this cell variant, experiments were designed that establish the functional significance of UDP-Xyl transport with respect to UDP-glucuronic acid production and xylosylation.  相似文献   

4.
5.
6.
Cell membranes predominantly consist of lamellar lipid bilayers. When studied in vitro, however, many membrane lipids can exhibit non-lamellar morphologies, often with cubic symmetries. An open issue is how lipid polymorphisms influence organelle and cell shape. Here, we used controlled dimerization of artificial membrane proteins in mammalian tissue culture cells to induce an expansion of the endoplasmic reticulum (ER) with cubic symmetry. Although this observation emphasizes ER architectural plasticity, we found that the changed ER membrane became sequestered into large autophagic vacuoles, positive for the autophagy protein LC3. Autophagy may be targeting irregular membrane shapes and/or aggregated protein. We suggest that membrane morphology can be controlled in cells.The observation that simple mixtures of amphiphilic (polar) lipids and water yield a rich flora of phase structures has opened a long-standing debate as to whether such membrane polymorphisms are relevant for living organisms (17). Lipid bilayers with planar geometry, termed lamellar symmetry, dominate the membrane structure of cells. However, this architecture comprises only a fraction of the structures seen with in vitro lipid-water systems (711). The propensity to form lamellar bilayers (a property exclusive to cylindrically shaped lipids) is flanked by a continuum of lipid structures that occur in a number of exotic and probably non-physiological non-bilayer configurations (3, 12). However, certain lipids, particularly those with smaller head groups and more bulky hydrocarbon chains, can adopt bilayered non-lamellar phases called cubic phases. Here the bilayer is curved everywhere in the form of saddle shapes corresponding to an energetically favorable minimal surface of zero mean curvature (1, 7). Because a substantial number of the lipids present in biological membranes, when studied as individual pure lipids, form cubic phases (13), cubic membranes have received particular interest in cell biology.Since the application of electron microscopy (EM)3 to the study of cell ultrastructure, unusual membrane morphologies have been reported for virtually every organelle (14, 15). However, interpretation of three-dimensional structures from two-dimensional electron micrographs is not easy (16). In seminal work, Landh (17) developed the method of direct template correlative matching, a technique that unequivocally assesses the presence of cubic membranes in biological specimens (16). Cubic phases adopt mathematically well defined three-dimensional configurations whose two-dimensional analogs have been derived (4, 17). In direct template correlative matching, electron micrographs are matched to these analogs. Cubic cell membrane geometries and in vitro cubic phases of purified lipid mixtures do differ in their lattice parameters; however, such deviations are thought to relate to differences in water activity and lipid to protein ratios (10, 14, 18). Direct template correlative matching has revealed thousands of examples of cellular cubic membranes in a broad survey of electron micrographs ranging from protozoa to human cells (14, 17) and, more recently, in the mitochondria of amoeba (19) and in subcellular membrane compartments associated with severe acute respiratory syndrome virus (20). Analysis of cellular cubic membranes has also been furthered by the development of EM tomography that confirmed the presence of cubic bilayers in the mitochondrial membranes of amoeba (21, 22).Although it is now clear that cubic membranes can exist in living cells, the generation of such architecture would appear tightly regulated, as evidenced by the dominance of lamellar bilayers in biology. In this light, we examined the capability and implications of generating cubic membranes in the endoplasmic reticulum (ER) of mammalian tissue culture cells. The ER is a spatially interconnected complex consisting of two domains, the nuclear envelope and the peripheral ER (2326). The nuclear envelope surrounds the nucleus and is composed of two continuous sheets of membranes, an inner and outer nuclear membrane connected to each other at nuclear pores. The peripheral ER constitutes a network of branching trijunctional tubules that are continuous with membrane sheet regions that occur in closer proximity to the nucleus. Recently it has been suggested that the classical morphological definition of rough ER (ribosome-studded) and smooth ER (ribosome-free) may correspond to sheet-like and tubular ER domains, respectively (27). The ER has a strong potential for cubic architectures, as demonstrated by the fact that the majority of cubic cell membranes in the EM record come from ER-derived structures (14, 17). Furthermore, ER cubic symmetries are an inducible class of organized smooth ER (OSER), a definition collectively referring to ordered smooth ER membranes (=stacked cisternae on the outer nuclear membrane, also called Karmelle (2830), packed sinusoidal ER (31), concentric membrane whorls (30, 3234), and arrays of crystalloid ER (3537)). Specifically, weak homotypic interactions between membrane proteins produce both a whorled and a sinusoidal OSER phenotype (38), the latter exhibiting a cubic symmetry (16, 39).We were able to produce OSER with cubic membrane morphology via induction of homo-dimerization of artificial membrane proteins. Interestingly, the resultant cubic membrane architecture was removed from the ER system by incorporation into large autophagic vacuoles. To assess whether these cubic symmetries were favored in the absence of cellular energy, we depleted ATP. To our surprise, the cells responded by forming large domains of tubulated membrane, suggesting that a cubic symmetry was not the preferred conformation of the system. Our results suggest that whereas the endoplasmic reticulum is capable of adopting cubic symmetries, both the inherent properties of the ER system and active cellular mechanisms, such as autophagy, can tightly control their appearance.  相似文献   

7.
In archaea and eukarya, box C/D ribonucleoprotein (RNP) complexes are responsible for 2′-O-methylation of tRNAs and rRNAs. The archaeal box C/D small RNP complex requires a small RNA component (sRNA) possessing Watson-Crick complementarity to the target RNA along with three proteins: L7Ae, Nop5p, and fibrillarin. Transfer of a methyl group from S-adenosylmethionine to the target RNA is performed by fibrillarin, which by itself has no affinity for the sRNA-target duplex. Instead, it is targeted to the site of methylation through association with Nop5p, which in turn binds to the L7Ae-sRNA complex. To understand how Nop5p serves as a bridge between the targeting and catalytic functions of the box C/D small RNP complex, we have employed alanine scanning to evaluate the interaction between the Pyrococcus horikoshii Nop5p domain and an L7Ae box C/D RNA complex. From these data, we were able to construct an isolated RNA-binding domain (Nop-RBD) that folds correctly as demonstrated by x-ray crystallography and binds to the L7Ae box C/D RNA complex with near wild type affinity. These data demonstrate that the Nop-RBD is an autonomously folding and functional module important for protein assembly in a number of complexes centered on the L7Ae-kinkturn RNP.Many biological RNAs require extensive modification to attain full functionality in the cell (1). Currently there are over 100 known RNA modification types ranging from small functional group substitutions to the addition of large multi-cyclic ring structures (2). Transfer RNA, one of many functional RNAs targeted for modification (3-6), possesses the greatest modification type diversity, many of which are important for proper biological function (7). Ribosomal RNA, on the other hand, contains predominantly two types of modified nucleotides: pseudouridine and 2′-O-methylribose (8). The crystal structures of the ribosome suggest that these modifications are important for proper folding (9, 10) and structural stabilization (11) in vivo as evidenced by their strong tendency to localize to regions associated with function (8, 12, 13). These roles have been verified biochemically in a number of cases (14), whereas newly emerging functional modifications are continually being investigated.Box C/D ribonucleoprotein (RNP)3 complexes serve as RNA-guided site-specific 2′-O-methyltransferases in both archaea and eukaryotes (15, 16) where they are referred to as small RNP complexes and small nucleolar RNPs, respectively. Target RNA pairs with the sRNA guide sequence and is methylated at the 2′-hydroxyl group of the nucleotide five bases upstream of either the D or D′ box motif of the sRNA (Fig. 1, star) (17, 18). In archaea, the internal C′ and D′ motifs generally conform to a box C/D consensus sequence (19), and each sRNA contains two guide regions ∼12 nucleotides in length (20). The bipartite architecture of the RNP potentially enables the complex to methylate two distinct RNA targets (21) and has been shown to be essential for site-specific methylation (22).Open in a separate windowFIGURE 1.Organization of the archaeal box C/D complex. The protein components of this RNP are L7Ae, Nop5p, and fibrillarin, which together bind a box C/D sRNA. The regions of the Box C/D sRNA corresponding to the conserved C, D, C′, and D′ boxes are labeled. The target RNA binds the sRNA through Watson-Crick pairing and is methylated by fibrillarin at the fifth nucleotide from the D/D′ boxes (star).In addition to the sRNA, the archaeal box C/D complex requires three proteins for activity (23): the ribosomal protein L7Ae (24, 25), fibrillarin, and the Nop56/Nop58 homolog Nop5p (Fig. 1). L7Ae binds to both box C/D and the C′/D′ motifs (26), which respectively comprise kink-turn (27) or k-loop structures (28), to initiate the assembly of the RNP (29, 30). Fibrillarin performs the methyl group transfer from the cofactor S-adenosylmethionine to the target RNA (31-33). For this to occur, the active site of fibrillarin must be positioned precisely over the specific 2′-hydroxyl group to be methylated. Although fibrillarin methylates this functional group in the context of a Watson-Crick base-paired helix (guide/target), it has little to no binding affinity for double-stranded RNA or for the L7Ae-sRNA complex (22, 26, 33, 34). Nop5p serves as an intermediary protein bringing fibrillarin to the complex through its association with both the L7Ae-sRNA complex and fibrillarin (22). Along with its role as an intermediary between fibrillarin and the L7Ae-sRNA complex, Nop5p possesses other functions not yet fully understood. For example, Nop5p self-dimerizes through a coiled-coil domain (35) that in most archaea and eukaryotic homologs includes a small insertion sequence of unknown function (36, 37). However, dimerization and fibrillarin binding have been shown to be mutually exclusive in Methanocaldococcus jannaschii Nop5p, potentially because of the presence of this insertion sequence (36). Thus, whether Nop5p is a monomer or a dimer in the active RNP is still under debate.In this study, we focus our attention on the Nop5p protein to investigate its interaction with a L7Ae box C/D RNA complex because both the fibrillarin-Nop5p and the L7Ae box C/D RNA interfaces are known from crystal structures (29, 35, 38). Individual residues on the surface of a monomeric form of Nop5p (referred to as mNop5p) (22) were mutated to alanine, and the effect on binding affinity for a L7Ae box C/D motif RNA complex was assessed through the use of electrophoretic mobility shift assays. These data reveal that residues important for binding cluster within the highly conserved NOP domain (39, 40). To demonstrate that this domain is solely responsible for the affinity of Nop5p for the preassembled L7Ae box C/D RNA complex, we expressed and purified it in isolation from the full Nop5p protein. The isolated Nop-RBD domain binds to the L7Ae box C/D RNA complex with nearly wild type affinity, demonstrating that the Nop-RBD is truly an autonomously folding and functional module. Comparison of our data with the crystal structure of the homologous spliceosomal hPrp31-15.5K protein-U4 snRNA complex (41) suggests the adoption of a similar mode of binding, further supporting a crucial role for the NOP domain in RNP complex assembly.  相似文献   

8.
Thimet oligopeptidase (EC 3.4.24.15; EP24.15) is an intracellular enzyme that has been proposed to metabolize peptides within cells, thereby affecting antigen presentation and G protein-coupled receptor signal transduction. However, only a small number of intracellular substrates of EP24.15 have been reported previously. Here we have identified over 100 peptides in human embryonic kidney 293 (HEK293) cells that are derived from intracellular proteins; many but not all of these peptides are substrates or products of EP24.15. First, cellular peptides were extracted from HEK293 cells and incubated in vitro with purified EP24.15. Then the peptides were labeled with isotopic tags and analyzed by mass spectrometry to obtain quantitative data on the extent of cleavage. A related series of experiments tested the effect of overexpression of EP24.15 on the cellular levels of peptides in HEK293 cells. Finally, synthetic peptides that corresponded to 10 of the cellular peptides were incubated with purified EP24.15 in vitro, and the cleavage was monitored by high pressure liquid chromatography and mass spectrometry. Many of the EP24.15 substrates identified by these approaches are 9–11 amino acids in length, supporting the proposal that EP24.15 can function in the degradation of peptides that could be used for antigen presentation. However, EP24.15 also converts some peptides into products that are 8–10 amino acids, thus contributing to the formation of peptides for antigen presentation. In addition, the intracellular peptides described here are potential candidates to regulate protein interactions within cells.Intracellular protein turnover is a crucial step for cell functioning, and if this process is impaired, the elevated levels of aged proteins usually lead to the formation of intracellular insoluble aggregates that can cause severe pathologies (1). In mammalian cells, most proteins destined for degradation are initially tagged with a polyubiquitin chain in an energy-dependent process and then digested to small peptides by the 26 S proteasome, a large proteolytic complex involved in the regulation of cell division, gene expression, and other key processes (2, 3). In eukaryotes, 30–90% of newly synthesized proteins may be degraded by proteasomes within minutes of synthesis (3, 4). In addition to proteasomes, other extralysosomal proteolytic systems have been reported (5, 6). The proteasome cleaves proteins into peptides that are typically 2–20 amino acids in length (7). In most cases, these peptides are thought to be rapidly hydrolyzed into amino acids by aminopeptidases (810). However, some intracellular peptides escape complete degradation and are imported into the endoplasmic reticulum where they associate with major histocompatibility complex class I (MHC-I)3 molecules and traffic to the cell surface for presentation to the immune system (1012). Additionally, based on the fact that free peptides added to the intracellular milieu can regulate cellular functions mediated by protein interactions such as gene regulation, metabolism, cell signaling, and protein targeting (13, 14), intracellular peptides generated by proteasomes that escape degradation have been suggested to play a role in regulating protein interactions (15). Indeed, oligopeptides isolated from rat brain tissue using the catalytically inactive EP24.15 (EC 3.4.24.15) were introduced into Chinese hamster ovarian-S and HEK293 cells and were found capable of altering G protein-coupled receptor signal transduction (16). Moreover, EP24.15 overexpression itself changed both angiotensin II and isoproterenol signal transduction, suggesting a physiological function for its intracellular substrates/products (16).EP24.15 is a zinc-dependent peptidase of the metallopeptidase M3 family that contains the HEXXH motif (17). This enzyme was first described as a neuropeptide-degrading enzyme present in the soluble fraction of brain homogenates (18). Whereas EP24.15 can be secreted (19, 20), its predominant location in the cytosol and nucleus suggests that the primary function of this enzyme is not the extracellular degradation of neuropeptides and hormones (21, 22). EP24.15 was shown in vivo to participate in antigen presentation through MHC-I (2325) and in vitro to bind (26) or degrade (27) some MHC-I associated peptides. EP24.15 has also been shown in vitro to degrade peptides containing 5–17 amino acids produced after proteasome digestion of β-casein (28). EP24.15 shows substrate size restriction to peptides containing from 5 to 17 amino acids because of its catalytic center that is located in a deep channel (29). Despite the size restriction, EP24.15 has a broad substrate specificity (30), probably because a significant portion of the enzyme-binding site is lined with potentially flexible loops that allow reorganization of the active site following substrate binding (29). Recently, it has also been suggested that certain substrates may be cleaved by an open form of EP24.15 (31). This characteristic is supported by the ability of EP24.15 to accommodate different amino acid residues at subsites S4 to S3′, which even includes the uncommon post-proline cleavage (30). Such biochemical and structural features make EP24.15 a versatile enzyme to degrade structurally unrelated oligopeptides.Previously, brain peptides that bound to catalytically inactive EP24.15 were isolated and identified using mass spectrometry (22). The majority of peptides captured by the inactive enzyme were intracellular protein fragments that efficiently interacted with EP24.15; the smallest peptide isolated in these assays contained 5 and the largest 17 amino acids (15, 16, 22, 32), which is within the size range previously reported for natural and synthetic substrates of EP24.15 (18, 30, 33, 34). Interestingly, the peptides released by the proteasome are in the same size range of EP24.15 competitive inhibitors/substrates (7, 35, 36). Taken altogether, these data suggest that in the intracellular environment EP24.15 could further cleave proteasome-generated peptides unrelated to MHC-I antigen presentation (15).Although the mutated inactive enzyme “capture” assay was successful in identifying several cellular protein fragments that were substrates for EP24.15, it also found some interacting peptides that were not substrates. In this study, we used several approaches to directly screen for cellular peptides that were cleaved by EP24.15. The first approach involved the extraction of cellular peptides from the HEK293 cell line, incubation in vitro with purified EP24.15, labeling with isotopic tags, and analysis by mass spectrometry to obtain quantitative data on the extent of cleavage. The second approach examined the effect of EP24.15 overexpression on the cellular levels of peptides in the HEK293 cell line. The third set of experiments tested synthetic peptides with purified EP24.15 in vitro, and examined cleavage by high pressure liquid chromatography and mass spectrometry. Collectively, these studies have identified a large number of intracellular peptides, including those that likely represent the endogenous substrates and products of EP24.15, and this original information contributes to a better understanding of the function of this enzyme in vivo.  相似文献   

9.
Frontotemporal lobar degeneration (FTLD) with inclusion body myopathy and Paget disease of bone is a rare, autosomal dominant disorder caused by mutations in the VCP (valosin-containing protein) gene. The disease is characterized neuropathologically by frontal and temporal lobar atrophy, neuron loss and gliosis, and ubiquitin-positive inclusions (FTLD-U), which are distinct from those seen in other sporadic and familial FTLD-U entities. The major component of the ubiquitinated inclusions of FTLD with VCP mutation is TDP-43 (TAR DNA-binding protein of 43 kDa). TDP-43 proteinopathy links sporadic amyotrophic lateral sclerosis, sporadic FTLD-U, and most familial forms of FTLD-U. Understanding the relationship between individual gene defects and pathologic TDP-43 will facilitate the characterization of the mechanisms leading to neurodegeneration. Using cell culture models, we have investigated the role of mutant VCP in intracellular trafficking, proteasomal function, and cell death and demonstrate that mutations in the VCP gene 1) alter localization of TDP-43 between the nucleus and cytosol, 2) decrease proteasome activity, 3) induce endoplasmic reticulum stress, 4) increase markers of apoptosis, and 5) impair cell viability. These results suggest that VCP mutation-induced neurodegeneration is mediated by several mechanisms.Frontotemporal lobar degeneration (FTLD)2 accounts for 10% of all late onset dementias and is the third most frequent neurodegenerative disease after Alzheimer disease and dementia with Lewy bodies (1). FTLD with ubiquitin-immunoreactive inclusions is genetically, clinically, and neuropathologically heterogeneous (2, 3). FTLD-U comprises several distinct entities, including sporadic forms and familial cases caused by mutations in the genes encoding VCP (valosin-containing protein), GRN (progranulin), CHMP2B (charged multivesicular body protein 2B), TDP-43 (TAR DNA-binding protein of 43 kDa) and an unknown gene linked to chromosome 9 (2, 3). Frontotemporal dementia with inclusion body myopathy and Paget disease of bone is a rare, autosomal dominant disorder caused by mutations in the VCP gene located on chromosome 9p13-p12 (4-10) (Fig. 1). This multisystem disease is characterized by progressive muscle weakness and atrophy, increased osteoclastic bone resorption, and early onset frontotemporal dementia, also called FTLD (9, 11). Mutations in VCP are also associated with dilatative cardiomyopathy with ubiquitin-positive inclusions (12). Neuropathologic features of FTLD with VCP mutation include frontal and temporal lobar atrophy, neuron loss and gliosis, and ubiquitin-positive inclusions (FTLD-U). The majority of aggregates are ubiquitin- and TDP-43-positive neuronal intranuclear inclusions (NIIs); a smaller proportion is made up of TDP-43-immunoreactive dystrophic neurites (DNs) and neuronal cytoplasmic inclusions (NCIs). A small number of inclusions are VCP-immunoreactive (5, 13). Pathologic TDP-43 in inclusions links a spectrum of diseases in which TDP-43 pathology is a primary feature, including FTLD-U, motor neuron disease, including amyotrophic lateral sclerosis, FTLD with motor neuron disease, and inclusion body myopathy and Paget disease of bone, as well as an expanding spectrum of other disorders in which TDP-43 pathology is secondary (14, 15).Open in a separate windowFIGURE 1.Model of pathogenic mutations and domains in valosin-containing protein. CDC48 (magenta), located within the N terminus (residues 22-108), binds the following cofactors: p47, gp78, and Npl4-Ufd1 (23-25, 28). There are two AAA-ATPase domains (AAA; blue) at residues 240-283 and 516-569, which are joined by two linker regions (L1 and L2; red).TDP-43 proteinopathy in FTLD with VCP mutation has a biochemical signature similar to that seen in other sporadic and familial cases of FTLD-U, including sporadic amyotrophic lateral sclerosis, FTLD-motor neuron disease, FTLD with progranulin (GRN) mutation, and FTLD linked to chromosome 9p (3, 16). TDP-43 proteinopathy in these disorders is characterized by hyperphosphorylation of TDP-43, ubiquitination, and cleavage to form C-terminal fragments detected only in insoluble brain extracts from affected brain regions (16). Identification of TDP-43 as the major component of the ubiquitin-immunoreactive inclusions of FTLD with VCP mutation supports the hypothesis that VCP gene mutations cause an alteration of VCP function, leading to TDP-43 proteinopathy.VCP/p97 (valosin-containing protein) is a member of the AAA (ATPase associated with diverse cellular activities) superfamily. The N-terminal domain of VCP has been shown to be involved in cofactor binding (CDC48 (cell division cycle protein 48)) and two AAA-ATPase domains that form a hexameric complex (Fig. 1) (17). Recently, it has been shown that the N-terminal domain of VCP binds phosphoinositides (18, 19). AKT (activated serine-threonine protein kinase) phosphorylates VCP and is required for constitutive VCP function (20, 21). AKT is activated through phospholipid binding and phosphorylation via the phosphoinositide 3-kinase signaling pathway, which is involved in cell survival (22). The lipid binding domain may recruit VCP to the cell membrane where it is phosphorylated by AKT (19).The diversity of VCP functions is modulated, in part, by a variety of intracellular cofactors, including p47, gp78, and Npl4-Ufd1 (23). Cofactor p47 has been shown to play a role in the maintenance and biogenesis of both the endoplasmic reticulum (ER) and Golgi apparatus (24). The structure of p47 contains a ubiquitin regulatory X domain that binds the N-terminus of VCP, and together they act as a chaperone to deliver membrane fusion machinery to the site of adjacent membranes (25). The function of the p47-VCP complex is dependent upon cell division cycle 2 (CDC2) serine-threonine kinase phosphorylation of p47 (26, 27). Also, VCP has been found to interact with the cytosolic tail of gp78, an ER membrane-spanning E3 ubiquitin ligase that exclusively binds VCP and enhances ER-associated degradation (ERAD) (28). The Npl4-Ufd1-VCP complex is involved in nuclear envelope assembly and targeting of proteins through the ubiquitin-proteasome system (29, 30). The cell survival response of this complex has been found to be important in DNA damage repair though activation by phosphorylation and its recruitment to double-stranded breaks (20, 31). The Npl4-Ufd1-VCP cytosolic complex is also recruited to the ER membrane, interacting with Derlin 1, VCP-interacting membrane proteins (VIMP), and other complexes. At the ER membrane, these misfolded proteins are targeted to the proteasome via ERAD (32-34). VCP also targets IKKβ for ubiquitination to the ubiquitin-proteasome system, implicating VCP in the cell survival pathway and neuroprotection (21, 35-37).To investigate the mechanism of neurodegeneration caused by VCP mutations, we first tested the hypothesis that VCP mutations decrease cell viability in vitro using a neuroblastoma SHSY-5Y cell line and then investigated cellular pathways that are known to lead to neurodegeneration, including decrease in proteasome activity, caspase-mediated degeneration, and a change in cellular localization of TDP-43.  相似文献   

10.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

11.
Rheb G-protein plays critical roles in the TSC/Rheb/mTOR signaling pathway by activating mTORC1. The activation of mTORC1 by Rheb can be faithfully reproduced in vitro by using mTORC1 immunoprecipitated by the use of anti-raptor antibody from mammalian cells starved for nutrients. The low in vitro kinase activity against 4E-BP1 of this mTORC1 preparation is dramatically increased by the addition of recombinant Rheb. On the other hand, the addition of Rheb does not activate mTORC2 immunoprecipitated from mammalian cells by the use of anti-rictor antibody. The activation of mTORC1 is specific to Rheb, because other G-proteins such as KRas, RalA/B, and Cdc42 did not activate mTORC1. Both Rheb1 and Rheb2 activate mTORC1. In addition, the activation is dependent on the presence of bound GTP. We also find that the effector domain of Rheb is required for the mTORC1 activation. FKBP38, a recently proposed mediator of Rheb action, appears not to be involved in the Rheb-dependent activation of mTORC1 in vitro, because the preparation of mTORC1 that is devoid of FKBP38 is still activated by Rheb. The addition of Rheb results in a significant increase of binding of the substrate protein 4E-BP1 to mTORC1. PRAS40, a TOR signaling (TOS) motif-containing protein that competes with the binding of 4EBP1 to mTORC1, inhibits Rheb-induced activation of mTORC1. A preparation of mTORC1 that is devoid of raptor is not activated by Rheb. Rheb does not induce autophosphorylation of mTOR. These results suggest that Rheb induces alteration in the binding of 4E-BP1 with mTORC1 to regulate mTORC1 activation.Rheb defines a unique member of the Ras superfamily G-proteins (1). We have shown that Rheb proteins are conserved and are found from yeast to human (2). Although yeast and fruit fly have one Rheb, mouse and human have two Rheb proteins termed Rheb1 (or simply Rheb) and Rheb2 (RhebL1) (2). Structurally, these proteins contain G1-G5 boxes, short stretches of amino acids that define the function of the Ras superfamily G-proteins including guanine nucleotide binding (1, 3, 4). Rheb proteins have a conserved arginine at residue 15 that corresponds to residue 12 of Ras (1). The effector domain required for the binding with downstream effectors encompasses the G2 box and its adjacent sequences (1, 5). Structural analysis by x-ray crystallography further shows that the effector domain is exposed to solvent, is located close to the phosphates of GTP especially at residues 35–38, and undergoes conformational change during GTP/GDP exchange (6). In addition, all Rheb proteins end with the CAAX (C is cysteine, A is an aliphatic amino acid, and X is the C-terminal amino acid) motif that signals farnesylation. In fact, we as well as others have shown that these proteins are farnesylated (79).Rheb plays critical roles in the TSC/Rheb/mTOR signaling, a signaling pathway that plays central roles in regulating protein synthesis and growth in response to nutrient, energy, and growth conditions (1014). Rheb is down-regulated by a TSC1·TSC2 complex that acts as a GTPase-activating protein for Rheb (1519). Recent studies established that the GAP domain of TSC2 defines the functional domain for the down-regulation of Rheb (20). Mutations in the Tsc1 or Tsc2 gene lead to tuberous sclerosis whose symptoms include the appearance of benign tumors called hamartomas at different parts of the body as well as neurological symptoms (21, 22). Overexpression of Rheb results in constitutive activation of mTOR even in the absence of nutrients (15, 16). Two mTOR complexes, mTORC1 and mTORC2, have been identified (23, 24). Whereas mTORC1 is involved in protein synthesis activation mediated by S6K and 4EBP1, mTORC2 is involved in the phosphorylation of Akt in response to insulin. It has been suggested that Rheb is involved in the activation of mTORC1 but not mTORC2 (25).Although Rheb is clearly involved in the activation of mTOR, the mechanism of activation has not been established. We as well as others have suggested a model that involves the interaction of Rheb with the TOR complex (2628). Rheb activation of mTOR kinase activity using immunoprecipitated mTORC1 was reported (29). Rheb has been shown to interact with mTOR (27, 30), and this may involve direct interaction of Rheb with the kinase domain of mTOR (27). However, this Rheb/mTOR interaction is a weak interaction and is not dependent on the presence of GTP bound to Rheb (27, 28). Recently, a different model proposing that FKBP38 (FK506-binding protein 38) mediates the activation of mTORC1 by Rheb was proposed (31, 32). In this model, FKBP38 binds mTOR and negatively regulates mTOR activity, and this negative regulation is blocked by the binding of Rheb to FKBP38. However, recent reports dispute this idea (33).To further characterize Rheb activation of mTOR, we have utilized an in vitro system that reproduces activation of mTORC1 by the addition of recombinant Rheb. We used mTORC1 immunoprecipitated from nutrient-starved cells using anti-raptor antibody and have shown that its kinase activity against 4E-BP1 is dramatically increased by the addition of recombinant Rheb. Importantly, the activation of mTORC1 is specific to Rheb and is dependent on the presence of bound GTP as well as an intact effector domain. FKBP38 is not detected in our preparation and further investigation suggests that FKBP38 is not an essential component for the activation of mTORC1 by Rheb. Our study revealed that Rheb enhances the binding of a substrate 4E-BP1 with mTORC1 rather than increasing the kinase activity of mTOR.  相似文献   

12.
13.
14.
Caprazamycins are potent anti-mycobacterial liponucleoside antibiotics isolated from Streptomyces sp. MK730-62F2 and belong to the translocase I inhibitor family. Their complex structure is derived from 5′-(β-O-aminoribosyl)-glycyluridine and comprises a unique N-methyldiazepanone ring. The biosynthetic gene cluster has been identified, cloned, and sequenced, representing the first gene cluster of a translocase I inhibitor. Sequence analysis revealed the presence of 23 open reading frames putatively involved in export, resistance, regulation, and biosynthesis of the caprazamycins. Heterologous expression of the gene cluster in Streptomyces coelicolor M512 led to the production of non-glycosylated bioactive caprazamycin derivatives. A set of gene deletions validated the boundaries of the cluster and inactivation of cpz21 resulted in the accumulation of novel simplified liponucleoside antibiotics that lack the 3-methylglutaryl moiety. Therefore, Cpz21 is assigned to act as an acyltransferase in caprazamycin biosynthesis. In vivo and in silico analysis of the caprazamycin biosynthetic gene cluster allows a first proposal of the biosynthetic pathway and provides insights into the biosynthesis of related uridyl-antibiotics.Caprazamycins (CPZs)2 (Fig. 1, 1) are liponucleoside antibiotics isolated from a fermentation broth of Streptomyces sp. MK730-62F2 (1, 2). They show excellent activity in vitro against Gram-positive bacteria, in particular against the genus Mycobacterium including Mycobacterium intracellulare, Mycobacterium avium, and Mycobacterium tuberculosis (3). In a pulmonary mouse model with M. tuberculosis H37Rv, administration of caprazamycin B exhibited a therapeutic effect but no significant toxicity (4). Structural elucidation (2) revealed a complex and unique composition of elements the CPZs share only with the closely related liposidomycins (LPMs, 2) (5). The core skeleton is the (+)-caprazol (5) composed of an N-alkylated 5′-(β-O-aminoribosyl)-glycyluridine, also known from FR-900493 (6) (6) and the muraymycins (7) (7), which is cyclized to form a rare diazepanone ring. Attached to the 3′″-OH are β-hydroxy fatty acids of different chain length resulting in CPZs A–G (1). They differ from the LPMs in the absence of a sulfate group at the 2″-position of the aminoribose and the presence of a permethylated l-rhamnose β-glycosidically linked to the 3-methylglutaryl (3-MG) moiety.Open in a separate windowFIGURE 1.Nucleoside antibiotics of the translocase I inhibitor family.The LPMs have been shown to inhibit biosynthesis of the bacterial cell wall by targeting the formation of lipid I (8). The CPZs are expected to act in the same way and are assigned to the growing number of translocase I inhibitors that include other nucleoside antibiotics, like the tunicamycins and mureidomycins (9). During peptidoglycan formation, translocase I catalyzes the transfer of UDP-MurNAc-pentapeptide to the undecaprenyl phosphate carrier to generate lipid I (10). This reaction is considered an unexploited and promising target for new anti-infective drugs (11).Recent investigations indicate that the 3″-OH group (12), the amino group of the aminoribosyl-glycyluridine, and an intact uracil moiety (13) are essential for the inhibition of the Escherichia coli translocase I MraY. The chemical synthesis of the (+)-caprazol (5) was recently accomplished (14), however, this compound only shows weak antibacterial activity. In contrast, the acylated compounds 3 and 4 exhibit strong growth inhibition of mycobacteria, suggesting a potential role of the fatty acid side chain in penetration of the bacterial cell (15, 16). Apparently, the acyl-caprazols (4) represent the most simplified antibiotically active liponucleosides and a good starting point for further optimization of this class of potential therapeutics.Although chemical synthesis and biological activity of CPZs and LPMs has been studied in some detail, their biosynthesis remains speculative and only few data exists about the formation of other translocase I inhibitors (17, 18). Nevertheless, we assume that the CPZ biosynthetic pathway is partially similar to that of LPMs, FR-90043 (6), and muraymycins (7) and presents a model for the comprehension and manipulation of liponucleoside formation. Considering the unique structural features of the CPZs we also expect some unusual biotransformations to be involved in the formation of, e.g. the (+)-caprazol.Here we report the identification and analysis of the CPZ gene cluster, the first cluster of a translocase I inhibitor. A set of gene disruption experiments provide insights into the biosynthetic origin of the CPZs and moreover, heterologous expression of the gene cluster allows the generation of novel bioactive derivatives by pathway engineering.  相似文献   

15.
The ApbC protein has been shown previously to bind and rapidly transfer iron-sulfur ([Fe-S]) clusters to an apoprotein (Boyd, J. M., Pierik, A. J., Netz, D. J., Lill, R., and Downs, D. M. (2008) Biochemistry 47, 8195–8202. This study utilized both in vivo and in vitro assays to examine the function of variant ApbC proteins. The in vivo assays assessed the ability of ApbC proteins to function in pathways with low and high demand for [Fe-S] cluster proteins. Variant ApbC proteins were purified and assayed for the ability to hydrolyze ATP, bind [Fe-S] cluster, and transfer [Fe-S] cluster. This study details the first kinetic analysis of ATP hydrolysis for a member of the ParA subfamily of “deviant” Walker A proteins. Moreover, this study details the first functional analysis of mutant variants of the ever expanding family of ApbC/Nbp35 [Fe-S] cluster biosynthetic proteins. The results herein show that ApbC protein needs ATPase activity and the ability to bind and rapidly transfer [Fe-S] clusters for in vivo function.Proteins containing iron-sulfur ([Fe-S]) clusters are employed in a wide array of metabolic functions (reviewed in Ref. 1). Research addressing the biosynthesis of the iron-molybdenum cofactor of nitrogenase in Azotobacter vinelandii led to the discovery of an operon (iscAnifnifUSVcysE1) involved in the biosynthesis of [Fe-S] clusters (reviewed in Ref. 2). Subsequent experiments led to the finding of two more systems involved in the de novo biosynthesis of [Fe-S] clusters, the isc and the suf systems (3, 4). Like Escherichia coli, the genome of Salmonella enterica serovar Typhimurium encodes for the isc and suf [Fe-S] cluster biosynthesis machinery.Recent studies have identified a number of additional or non-isc/-suf-encoded proteins that are involved in bacterial [Fe-S] cluster biosynthesis and repair. Examples include the following: CyaY, an iron-binding protein believed to be involved in iron trafficking and iron delivery (57); YggX, an Fe2+-binding protein that protects the cell from oxidative stress (8, 9); ErpA, an alternate A-type [Fe-S] cluster scaffolding protein (10); NfuA, a proposed intermediate [Fe-S] delivery protein (1113); YtfE, a protein proposed to be involved in [Fe-S] cluster repair (14, 15); and CsdA-CsdE, an alternative cysteine desulferase (16).Analysis of the metabolic network anchored to thiamine biosynthesis in S. enterica identified lesions in three non-isc or -suf loci that compromise Fe-S metabolism as follows: apbC, apbE, and rseC (1721). This metabolic system was subsequently used to dissect a role for cyaY and gshA in [Fe-S] cluster metabolism (6, 22, 23). Of these, the apbC (mrp in E. coli) locus was identified as the predominant site of lesions that altered thiamine synthesis by disrupting [Fe-S] cluster metabolism (17, 18).ApbC is a member of the ParA subfamily of proteins that have a wide array of functions, including electron transfer (24), initiation of cell division (25), and DNA segregation (26, 27). Importantly, ATP hydrolysis is required for function of all well characterized members of this subfamily, and all members contain a “deviant” Walker A motif, which contains two lysine residues instead of one (GKXXXGK(S/T)) (28). ApbC has been shown to hydrolyze ATP (17).Recently, five proteins with a high degree of identity to ApbC have been shown to be involved in [Fe-S] cluster metabolism in eukaryotes. The sequence alignments of the central portion of these proteins and bacterial ApbC are shown in Fig. 1. HCF101 was demonstrated to be involved in chloroplast [Fe-S] cluster metabolism (29, 30). The CFD1, Npb35, and huNbp35 (formally Nubp1) proteins were demonstrated to be involved in cytoplasmic [Fe-S] cluster metabolism (31, 32). Ind1 was demonstrated to be involved in the maturation of [Fe-S] clusters in the mitochondrial enzyme NADH:ubiquinone oxidoreductase (33). There is currently no report of any of these proteins hydrolyzing ATP.Open in a separate windowFIGURE 1.Protein sequence alignments of members of the ApbC/Nbp35 subfamily of ParA family of proteins. Protein alignments were assembled using the Clustal_W method in the Lasergene® software and show only the central portion of the proteins, which have the highest sequence conservation. The three boxed areas highlight the Walker A box, conserved Ser residue, and CXXC motif. Proteins listed are as follows: ApbC (S. enterica serovar Typhimurium LT2), CFD1 (S. cerevisiae), Nbp35 (S. cerevisiae), HCF101 (Arabidopsis thaliana), huNpb35 (formally Nubp1) (Homo sapiens), and Ind1 (Candida albicans).Biochemical analysis of ApbC indicated that it could bind and transfer [Fe-S] clusters to Saccharomyces cerevisiae apo-isopropylmalate isomerase (34). Additional genetic studies indicated that ApbC has a degree of functional redundancy with IscU, a known [Fe-S] cluster scaffolding protein (35, 36).In this study we investigate the correlation between the biochemical properties of ApbC (i.e. ATPase activity, [Fe-S] cluster binding, and [Fe-S] cluster transfer rates) and the in vivo function of this protein. This is the first detailed kinetic analysis of ATP hydrolysis for a member of the ParA subfamily of deviant Walker A proteins and the first functional analysis of a member of the ever expanding family of ApbC/Nbp35 proteins. Data presented indicate that noncomplementing variants have distinct biochemical properties that place them in three distinct classes.  相似文献   

16.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

17.
The causative agent of Legionnaires disease, Legionella pneumophila, forms a replicative vacuole in phagocytes by means of the intracellular multiplication/defective organelle trafficking (Icm/Dot) type IV secretion system and translocated effector proteins, some of which subvert host GTP and phosphoinositide (PI) metabolism. The Icm/Dot substrate SidC anchors to the membrane of Legionella-containing vacuoles (LCVs) by specifically binding to phosphatidylinositol 4-phosphate (PtdIns(4)P). Using a nonbiased screen for novel L. pneumophila PI-binding proteins, we identified the Rab1 guanine nucleotide exchange factor (GEF) SidM/DrrA as the predominant PtdIns(4)P-binding protein. Purified SidM specifically and directly bound to PtdIns(4)P, whereas the SidM-interacting Icm/Dot substrate LidA preferentially bound PtdIns(3)P but also PtdIns(4)P, and the L. pneumophila Arf1 GEF RalF did not bind to any PIs. The PtdIns(4)P-binding domain of SidM was mapped to the 12-kDa C-terminal sequence, termed “P4M” (PtdIns4P binding of SidM/DrrA). The isolated P4M domain is largely helical and displayed higher PtdIns(4)P binding activity in the context of the α-helical, monomeric full-length protein. SidM constructs containing P4M were translocated by Icm/Dot-proficient L. pneumophila and localized to the LCV membrane, indicating that SidM anchors to PtdIns(4)P on LCVs via its P4M domain. An L. pneumophila ΔsidM mutant strain displayed significantly higher amounts of SidC on LCVs, suggesting that SidM and SidC compete for limiting amounts of PtdIns(4)P on the vacuole. Finally, RNA interference revealed that PtdIns(4)P on LCVs is specifically formed by host PtdIns 4-kinase IIIβ. Thus, L. pneumophila exploits PtdIns(4)P produced by PtdIns 4-kinase IIIβ to anchor the effectors SidC and SidM to LCVs.The Gram-negative pathogen Legionella pneumophila is the causative agent of Legionnaires disease, but it evolved as a parasite of various species of environmental predatory protozoa, including the social amoeba Dictyostelium discoideum (1, 2). The human disease is linked to the inhalation of contaminated aerosols, followed by replication in alveolar macrophages. To accommodate the transfer between host cells, L. pneumophila alternates between replicative and transmissive phases, the regulation of which includes an apparent quorum-sensing system (35).In macrophages and amoebae, L. pneumophila forms a replicative compartment, the Legionella-containing vacuole (LCV).3 LCVs avoid fusion with lysosomes (6), intercept vesicular traffic at endoplasmic reticulum (ER) exit sites (7), and fuse with the ER (810). The uptake of L. pneumophila and formation of LCVs in macrophages and amoebae depends on the Icm/Dot type IV secretion system (T4SS) (1114). Although more than 100 Icm/Dot substrates (“effector” proteins) have been identified to date, only few are functionally characterized, including effectors that interfere with host cell signal transduction, vesicle trafficking, or apoptotic pathways (1518).Two Icm/Dot-translocated substrates, SidM/DrrA (19, 20) and RalF (21), have been characterized as guanine nucleotide exchange factors (GEFs) for the Rho subfamily of small GTPases. These bacterial GEFs are recruited to and activate their targets on LCVs. Small GTPases of the Rho subfamily are involved in many eukaryotic signal transduction pathways and in actin cytoskeleton regulation (22). Inactive Rho GTPases bind GDP and a guanine nucleotide dissociation inhibitor (GDI). The GTPases are activated by removal of the GDI and the exchange of GDP with GTP by GEFs, which promotes the interaction with downstream effector proteins, such as protein or lipid kinases and various adaptor proteins. The cycle is closed by hydrolysis of the bound GTP, which is mediated by GTPase-activating proteins.SidM is a GEF for Rab1, which is essential for ER to Golgi vesicle transport, and additionally, SidM acts as a GDI displacement factor (GDF) to activate Rab1 (23, 24). The function of SidM is assisted by the Icm/Dot substrate LidA, which also localizes to LCVs. LidA preferentially binds to activated Rab1, thus supporting the recruitment of early secretory vesicles by SidM (19, 20, 23, 25, 26). Another Icm/Dot substrate, LepB (27), contributes to Rab1-mediated membrane cycling by inactivating Rab1 through its GTPase-activating protein function, thus acting as an antagonist of SidM (24).The Icm/Dot substrate RalF recruits and activates the small GTPase ADP-ribosylation factor 1 (Arf1), which is involved in retrograde vesicle transport from Golgi to ER (21). Dominant negative Arf1 (7, 28) or knockdown of Arf1 by RNA interference (29) impairs the formation of LCVs, as well as the recruitment of the Icm/Dot substrate SidC to the LCV (30).SidC and its paralogue SdcA localize to the LCV membrane (31), where the proteins specifically bind to the host cell lipid phosphatidylinositol 4-phosphate (PtdIns(4)P) (32, 33). Phosphoinositides (PIs) regulate eukaryotic receptor-mediated signal transduction, actin remodeling, and membrane dynamics (34, 35). PtdIns(4)P is present on the cytoplasmic membrane, but localizes preferentially to the trans-Golgi network (TGN), where this PI is produced by an Arf-dependent recruitment of PtdIns(4)P kinase IIIβ (PI4K IIIβ) (36) to promote trafficking along the secretory pathway. Recently, PtdIns(4)P was found to also mediate the export of early secretory vesicles from ER exit sites (37). At present, the L. pneumophila effector proteins that mediate exploitation of host PI signaling remain ill defined.In a nonbiased screen for L. pneumophila PI-binding proteins using different PIs coupled to agarose beads, we identified SidM as a major PtdIns(4)P-binding effector. We mapped its PtdIns(4)P binding activity to a novel P4M domain within a 12-kDa C-terminal sequence. SidM constructs, including the P4M domain, were found to be translocated and bind the LCV membrane, where the levels of PtdIns(4)P are controlled by PI4K IIIβ.  相似文献   

18.
19.
20.
ATP-binding cassette (ABC) transporters transduce the free energy of ATP hydrolysis to power the mechanical work of substrate translocation across cell membranes. MsbA is an ABC transporter implicated in trafficking lipid A across the inner membrane of Escherichia coli. It has sequence similarity and overlapping substrate specificity with multidrug ABC transporters that export cytotoxic molecules in humans and prokaryotes. Despite rapid advances in structure determination of ABC efflux transporters, little is known regarding the location of substrate-binding sites in the transmembrane segment and the translocation pathway across the membrane. In this study, we have mapped residues proximal to the daunorubicin (DNR)-binding site in MsbA using site-specific, ATP-dependent quenching of DNR intrinsic fluorescence by spin labels. In the nucleotide-free MsbA intermediate, DNR-binding residues cluster at the cytoplasmic end of helices 3 and 6 at a site accessible from the membrane/water interface and extending into an aqueous chamber formed at the interface between the two transmembrane domains. Binding of a nonhydrolyzable ATP analog inverts the transporter to an outward-facing conformation and relieves DNR quenching by spin labels suggesting DNR exclusion from proximity to the spin labels. The simplest model consistent with our data has DNR entering near an elbow helix parallel to the water/membrane interface, partitioning into the open chamber, and then translocating toward the periplasm upon ATP binding.ATP-binding cassette (ABC)2 transporters transduce the energy of ATP hydrolysis to power the movement of a wide range of substrates across the cell membranes (1, 2). They constitute the largest family of prokaryotic transporters, import essential cell nutrients, flip lipids, and export toxic molecules (3). Forty eight human ABC transporters have been identified, including ABCB1, or P-glycoprotein, which is implicated in cross-resistance to drugs and cytotoxic molecules (4, 5). Inherited mutations in these proteins are linked to diseases such as cystic fibrosis, persistent hypoglycemia of infancy, and immune deficiency (6).The functional unit of an ABC transporter consists of four modules. Two highly conserved ABCs or nucleotide-binding domains (NBDs) bind and hydrolyze ATP to supply the active energy for transport (7). ABCs drive the mechanical work of proteins with diverse functions ranging from membrane transport to DNA repair (3, 5). Substrate specificity is determined by two transmembrane domains (TMDs) that also provide the translocation pathway across the bilayer (7). Bacterial ABC exporters are expressed as monomers, each consisting of one NBD and one TMD, that dimerize to form the active transporter (3). The number of transmembrane helices and their organization differ significantly between ABC importers and exporters reflecting the divergent structural and chemical nature of their substrates (1, 8, 9). Furthermore, ABC exporters bind substrates directly from the cytoplasm or bilayer inner leaflet and release them to the periplasm or bilayer outer leaflet (10, 11). In contrast, bacterial importers have their substrates delivered to the TMD by a dedicated high affinity substrate-binding protein (12).In Gram-negative bacteria, lipid A trafficking from its synthesis site on the inner membrane to its final destination in the outer membrane requires the ABC transporter MsbA (13). Although MsbA has not been directly shown to transport lipid A, suppression of MsbA activity leads to cytoplasmic accumulation of lipid A and inhibits bacterial growth strongly suggesting a role in translocation (14-16). In addition to this role in lipid A transport, MsbA shares sequence similarity with multidrug ABC transporters such as human ABCB1, LmrA of Lactococcus lactis, and Sav1866 of Staphylococcus aureus (16-19). ABCB1, a prototype of the ABC family, is a plasma membrane protein whose overexpression provides resistance to chemotherapeutic agents in cancer cells (1). LmrA and MsbA have overlapping substrate specificity with ABCB1 suggesting that both proteins can function as drug exporters (18, 20). Indeed, cells expressing MsbA confer resistance to erythromycin and ethidium bromide (21). MsbA can be photolabeled with the ABCB1/LmrA substrate azidopine and can transport Hoechst 33342 (H33342) across membrane vesicles in an energy-dependent manner (21).The structural mechanics of ABC exporters was revealed from comparison of the MsbA crystal structures in the apo- and nucleotide-bound states as well as from analysis by spin labeling EPR spectroscopy in liposomes (17, 19, 22, 23). The energy harnessed from ATP binding and hydrolysis drives a cycle of NBD association and dissociation that is transmitted to induce reorientation of the TMD from an inward- to outward-facing conformation (17, 19, 22). Large amplitude motion closes the cytoplasmic end of a chamber found at the interface between the two TMDs and opens it to the periplasm (23). These rearrangements lead to significant changes in chamber hydration, which may drive substrate translocation (22).Substrate binding must precede energy input, otherwise the cycle is futile, wasting the energy of ATP hydrolysis without substrate extrusion (7). Consistent with this model, ATP binding reduces ABCB1 substrate affinity, potentially through binding site occlusion (24-26). Furthermore, the TMD substrate-binding event signals the NBD to stimulate ATP hydrolysis increasing transport efficiency (1, 27, 28). However, there is a paucity of information regarding the location of substrate binding, the transport pathway, and the structural basis of substrate recognition by ABC exporters. In vitro studies of MsbA substrate specificity identify a broad range of substrates that stimulate ATPase activity (29). In addition to the putative physiological substrates lipid A and lipopolysaccharide (LPS), the ABCB1 substrates Ilmofosine, H33342, and verapamil differentially enhance ATP hydrolysis of MsbA (29, 30). Intrinsic MsbA tryptophan (Trp) fluorescence quenching by these putative substrate molecules provides further support of interaction (29).Extensive biochemical analysis of ABCB1 and LmrA provides a general model of substrate binding to ABC efflux exporters. This so-called “hydrophobic cleaner model” describes substrates binding from the inner leaflet of the bilayer and then translocating through the TMD (10, 31, 32). These studies also identified a large number of residues involved in substrate binding and selectivity (33). When these crucial residues are mapped onto the crystal structures of MsbA, a subset of homologous residues clusters to helices 3 and 6 lining the putative substrate pathway (34). Consistent with a role in substrate binding and specificity, simultaneous replacement of two serines (Ser-289 and Ser-290) in helix 6 of MsbA reduces binding and transport of ethidium and taxol, although H33342 and erythromycin interactions remain unaffected (34).The tendency of lipophilic substrates to partition into membranes confounds direct analysis of substrate interactions with ABC exporters (35, 36). Such partitioning may promote dynamic collisions with exposed Trp residues and nonspecific cross-linking in photo-affinity labeling experiments. In this study, we utilize a site-specific quenching approach to identify residues in the vicinity of the daunorubicin (DNR)-binding site (37). Although the data on DNR stimulation of ATP hydrolysis is inconclusive (20, 29, 30), the quenching of MsbA Trp fluorescence suggests a specific interaction. Spin labels were introduced along transmembrane helices 3, 4, and 6 of MsbA to assess their ATP-dependent quenching of DNR fluorescence. Residues that quench DNR cluster along the cytoplasmic end of helices 3 and 6 consistent with specific binding of DNR. Furthermore, many of these residues are not lipid-exposed but face the putative substrate chamber formed between the two TMDs. These residues are proximal to two Trps, which likely explains the previously reported quenching (29). Our results suggest DNR partitions to the membrane and then binds MsbA in a manner consistent with the hydrophobic cleaner model. Interpretation in the context of the crystal structures of MsbA identifies a putative translocation pathway through the transmembrane segment.  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