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.
SEC35 was identified in a novel screen for temperature-sensitive mutants in the secretory pathway of the yeast Saccharomyces cerevisiae (Wuestehube et al., 1996. Genetics. 142:393–406). At the restrictive temperature, the sec35-1 strain exhibits a transport block between the ER and the Golgi apparatus and accumulates numerous vesicles. SEC35 encodes a novel cytosolic protein of 32 kD, peripherally associated with membranes. The temperature-sensitive phenotype of sec35-1 is efficiently suppressed by YPT1, which encodes the rab-like GTPase required early in the secretory pathway, or by SLY1-20, which encodes a dominant form of the ER to Golgi target -SNARE–associated protein Sly1p. Weaker suppression is evident upon overexpression of genes encoding the vesicle-SNAREs SEC22, BET1, or YKT6. The cold-sensitive lethality that results from deleting SEC35 is suppressed by YPT1 or SLY1-20. These genetic relationships suggest that Sec35p acts upstream of, or in conjunction with, Ypt1p and Sly1p as was previously found for Uso1p. Using a cell-free assay that measures distinct steps in vesicle transport from the ER to the Golgi, we find Sec35p is required for a vesicle docking stage catalyzed by Uso1p. These genetic and biochemical results suggest Sec35p acts with Uso1p to dock ER-derived vesicles to the Golgi complex.Protein transport through the secretory pathway occurs via transport vesicles under the direction of a large set of protein components (Rothman, 1994). The process can be divided into three stages: (a) vesicle budding, (b) vesicle docking, and (c) membrane fusion, with distinct sets of proteins mediating each phase. The budding step involves recruitment of coat proteins to the membrane and culminates with the release of coated vesicles (Schekman and Orci, 1996). The docking reaction is likely to require a set of integral membrane proteins on the vesicle and target membranes, termed v-SNAREs1 and t-SNAREs (vesicle- and target membrane-soluble N-ethylmaleimide–sensitive fusion protein [NSF] attachment protein [SNAP] receptors, respectively), that are thought to confer specificity through their pair-wise interactions (Söllner et al., 1993b ). Small GTP-binding proteins of the rab family also assist in the docking process (Ferro-Novick and Novick, 1993), but their precise function is not known. The fusion step ensues after docking and results in the delivery of the vesicular cargo to the next compartment in the secretory pathway.Vesicular transport from the ER to the Golgi apparatus in the yeast Saccharomyces cerevisiae has been extensively characterized. Transport vesicle budding involves the assembly of the COPII coat, composed of the Sec13p/Sec31p (Pryer et al., 1993; Salama et al., 1993; Barlowe et al., 1994) and Sec23p/Sec24p heterodimers (Hicke and Schekman, 1989; Hicke et al., 1992), under the direction of an integral membrane protein, Sec12p (Nakano et al., 1988; Barlowe and Schekman, 1993), a small GTPase, Sar1p (Nakano and Muramatsu, 1989), and a multidomain protein, Sec16p (Espenshade et al., 1995; Shaywitz et al., 1997). Docking is thought to require a tethering event mediated by Uso1p (Cao et al., 1998), the yeast homologue of mammalian p115 (Barroso et al., 1995; Sapperstein et al., 1995), followed by or concurrent with the interaction of a set of ER to Golgi v-SNAREs, Bet1p, Bos1p, Sec22p (Newman and Ferro-Novick, 1987; Newman et al., 1990; Ossig et al., 1991; Shim et al., 1991; Søgaard et al., 1994) and perhaps Ykt6p (Søgaard et al., 1994; McNew et al., 1997), with the cognate t-SNARE on the Golgi, Sed5p (Hardwick and Pelham, 1992). For some time it was thought that fusion may be initiated by disassembly of the v/t-SNARE complex (Söllner et al., 1993a ) by yeast SNAP, Sec17p, (Griff et al., 1992) and NSF, Sec18p (Eakle et al., 1988; Wilson et al., 1989). However, this concept has been challenged by studies with a yeast system that reconstitutes homotypic vacuolar fusion, which suggests the action of Sec18p is before vesicle docking (Mayer et al., 1996; Mayer and Wickner, 1997). In addition, a prefusion role for NSF has been supported by the recent finding that liposomes bearing SNAREs alone can fuse in the absence of NSF (Weber et al., 1998).Several proteins involved in the regulation of yeast ER to Golgi v/t-SNARE complex assembly have been identified, including Ypt1p, Uso1p, and Sly1p. Ypt1p is a member of the rab family of small GTP-binding proteins that have been identified as important components of almost every stage in the secretory pathway (Ferro-Novick and Novick, 1993). Hydrolysis of GTP by rab-like proteins has been hypothesized to provide the regulatory switch that controls the fidelity of vesicular transport (Bourne, 1988). A second protein, Uso1p (Nakajima et al., 1991), appears to function in the same pathway as Ypt1p (Sapperstein et al., 1996), and both proteins have been demonstrated to be essential for SNARE complex assembly (Søgaard et al., 1994; Sapperstein et al., 1996; Lupashin and Waters, 1997). The third protein, Sly1p, is associated with the t-SNARE Sed5p (Søgaard et al., 1994). SLY1 is an essential gene in yeast (Dascher et al., 1991; Ossig et al., 1991), and Sly1p is required for ER to Golgi transport in vitro (Lupashin et al., 1996) and in vivo (Ossig et al., 1991). However, several lines of evidence, particularly from Sly1p homologues in other organisms, indicate that Sly1p may also function as a negative regulator of v/t-SNARE complex assembly, perhaps by preventing the association of the v- and t-SNAREs (Hosono et al., 1992; Pevsner et al., 1994; Schulze et al., 1994). A dominant allele of SLY1, termed SLY1-20, is capable of suppressing mutations in YPT1 and USO1, including complete deletions (Dascher et al., 1991; Sapperstein et al., 1996). Thus, in the presence of Sly1-20p, two components required for SNARE complex assembly are no longer essential. We have proposed a model (Sapperstein et al., 1996; Lupashin and Waters, 1997) in which Ypt1p and Uso1p function to relieve the inhibitory action of Sly1p on SNARE complex assembly. In this model Sly1-20p can be thought of as a noninhibitory form of SLY1 that renders Ypt1p and Uso1p superfluous.We believe that the ability of SLY1-20 to suppress defects in upstream docking regulators can be used to identify additional components involved in the regulation of vesicular docking. We have undertaken a genetic screen (to be presented elsewhere) to isolate novel components in this pathway which, when mutated, depend upon Sly1-20p for viability. In the course of this work, we discovered that two recently identified mutants, sec34 and sec35, can be suppressed by SLY1-20 and thus satisfy the criterion of our screen. These mutants were isolated in a novel screen to identify components involved in transport at any step between the ER and the trans-Golgi network (i.e., the Kex2p compartment) in yeast (Wuestehube et al., 1996). Both sec34 and sec35 accumulate the core-glycosylated form of secretory proteins at the nonpermissive temperature, indicating a block in ER to Golgi transport. Furthermore, electron microscopy indicated that both sec34 and sec35 accumulate numerous vesicles upon shift to the restrictive temperature (Wuestehube et al., 1996), a hallmark of genes whose protein products are involved in the docking or fusion phase of transport (Kaiser and Schekman, 1990). In this report we describe the cloning of SEC35 and analysis of its genetic interactions with other secretory genes. Strong genetic interaction between SEC35 and SLY1, YPT1, and USO1 suggests that Sec35p may function in vesicle docking. To test this possibility, we devised an in vitro transport assay that depends on the addition of purified Sec35p and Uso1p. Vesicles synthesized in the absence of functional Sec35p do not fuse with the Golgi compartment and remain as freely diffusible intermediates. Upon addition of Sec35p and Uso1p, vesicles dock to the Golgi and proceed to membrane fusion. Requirements for Sec35p at the vesicle docking step correlates our genetic experiments with the biochemically distinguishable steps of vesicle docking and membrane fusion.  相似文献   

2.
Glucose regulates the degradation of the key gluconeogenic enzyme, fructose-1,6-bisphosphatase (FBPase), in Saccharomyces cerevisiae. FBPase is targeted from the cytosol to a novel type of vesicle, and then to the vacuole for degradation when yeast cells are transferred from medium containing poor carbon sources to fresh glucose. To identify proteins involved in the FBPase degradation pathway, we cloned our first VID (vacuolar import and degradation) gene. The VID24 gene was identified by complementation of the FBPase degradation defect of the vid24-1 mutant. Vid24p is a novel protein of 41 kD and is synthesized in response to glucose. Vid24p is localized to the FBPase-containing vesicles as a peripheral membrane protein. In the absence of functional Vid24p, FBPase accumulates in the vesicles and fails to move to the vacuole, suggesting that Vid24p regulates FBPase targeting from the vesicles to the vacuole. FBPase sequestration into the vesicles is not affected in the vid24-1 mutant, indicating that Vid24p acts after FBPase sequestration into the vesicles has occurred. Vid24p is the first protein identified that marks the FBPase-containing vesicles and plays a critical role in delivering FBPase from the vesicles to the vacuole for degradation.Protein degradation is an important process that is tightly regulated. In mammalian cells, serum starvation induces protein degradation by lysosomes (Dice, 1990; Hayes and Dice, 1996). Cytosolic proteins containing a pentapeptide sequence are targeted to the lysosome for degradation in a process mediated by a heat shock protein (Chiang and Dice, 1988; Chiang et al., 1989; Terlecky et al., 1992; Terlecky and Dice, 1993; Cuervo et al., 1994). The receptor protein for this selective proteolysis pathway has been identified recently to be LGP96 (Cuervo and Dice, 1996). Overexpression of the receptor protein increases the degradation of cytosolic proteins in lysosomes both in vivo and in vitro (Cuervo and Dice, 1996).In Saccharomyces cerevisiae, the vacuole is functionally homologous to the lysosome and takes up proteins by several mechanisms. Most vacuole resident proteinases such as carboxypeptidase Y (CPY)1 enter the vacuole through the secretory pathway (Hasilik and Tanner, 1978; Hemmings et al., 1981; Rothman and Stevens, 1986; Banta et al., 1988; Jones, 1991). CPY is synthesized and processed sequentially in the ER and the Golgi. Sorting occurs in the late Golgi by the CPY receptor encoded by the PEP1/ VPS10 gene (Marcusson et al., 1994; Cooper and Stevens, 1996). CPY is delivered to the vacuole from the prevacuolar or endosomal compartment and the receptor protein recycles back to the Golgi (Marcusson et al., 1994; Cooper and Stevens, 1996). Other vacuolar proteins such as α-mannosidase or aminopeptidase I are imported from the cytosol to the vacuole, independent of the secretory pathway (Yoshihisa and Anraku, 1990; Klionsky et al., 1992; Harding et al., 1995, 1996; Scott et al., 1996). Plasma membrane proteins can be internalized by endocytosis and transported through early endosomes to late endosomes, from which they are directed to the vacuole for degradation (Davis et al., 1993; Raths et al., 1993; Kolling and Hollenberg, 1994; Schandel and Jennes, 1994; Lai et al., 1995; Riballo et al., 1995). Organelles such as peroxisomes or mitochondria can be engulfed by the vacuoles by autophagy (Takeshige et al., 1992; Tuttle and Dunn, 1995; Chiang et al., 1996). The key gluconeogenic enzyme, fructose-1,6-bisphosphatase (FBPase), is induced when Saccharomyces cerevisiae cells are grown in medium containing poor carbon sources. When cells are transferred to medium containing fresh glucose, FBPase is rapidly inactivated (Gancedo, 1971). Using isogenic strains differing only at the PEP4 gene, we have demonstrated that FBPase is targeted from the cytosol to the vacuole for degradation when cells are transferred from poor carbon sources to fresh glucose (Chiang and Schekman, 1991). The PEP4 gene encodes proteinase A, whose activity is required for the maturation of proteinase B and proteinase C (Zubenko and Jones, 1981; Jones, 1991). As a result, the pep4 strain reduces the vacuolar proteolytic activity to 30% of the wild-type level (Zubenko and Jones, 1981; Jones, 1991; Chiang et al., 1996). The glucose-induced distribution of FBPase from the cytosol to the vacuole has been observed in the pep4 cell by cell fractionation techniques, immunofluorescence microscopy, and immunoelectron microscopy (Chiang and Schekman, 1991; Chiang et al., 1996). FBPase targeting into the vacuole always occurs, regardless of whether cells are transferred to glucose from acetate, ethanol, galactose, or oleate (Chiang and Schekman, 1994; Chiang et al., 1996).To dissect the FBPase degradation pathway, we have taken a genetic approach. Several vid (vacuolar import and degradation) mutants that fail to degrade FBPase in response to glucose have been isolated (Hoffman and Chiang, 1996). Most vid mutants block FBPase in the cytosol. However, in the vid14-1, vid15-1, and vid16-1 mutants, FBPase is found in punctate structures in the cytoplasm. When cell extracts from one of these mutants are fractionated, a substantial amount of FBPase is found in the high speed pellet, suggesting that FBPase is associated with intracellular structures in these mutants (Hoffman and Chiang, 1996). This association is also observed in wild-type cells (Huang and Chiang, 1997).The FBPase-containing vesicles have been purified from wild-type cells to near homogeneity using a combination of differential centrifugation, gel filtration, and equilibrium centrifugation in sucrose gradients (Huang and Chiang, 1997). The purified fractions contain 30–40-nm-diam vesicles and are essentially free of other organelles. Kinetic studies indicate that FBPase association with these vesicles is induced by glucose, occurs only transiently, and precedes the association with the vacuole. The FBPase-containing vesicles are distinct from mitochondria, peroxisomes, endosomes, vacuoles, ER, Golgi, or transport vesicles such as the coat protein (COPI or COPII)-containing vesicles as analyzed by protein markers and electron microscopy (Huang and Chiang, 1997).The vesicles were predicted to contain proteins involved in FBPase targeting and sequestration into the vesicles, as well as proteins participating in carrying FBPase from the vesicles to the vacuole for degradation. To identify such factors, we cloned our first VID gene. The VID24 gene was identified by complementation of the degradation defect of the vid24-1 mutant. Vid24p is a novel 41-kD protein and is synthesized in response to glucose. A significant portion of the Vid24p is localized to the FBPase-containing vesicles as a peripheral protein. The deletion of Vid24p abolishes the degradation of FBPase, but does not cause significant change in growth, sporulation, germination, osmolarity sensitivity, or processing of CPY. In the absence of functional Vid24p, FBPase accumulates in the vesicles and fails to move to the vacuole. FBPase is sequestered inside the vesicles in the vid24-1 mutant, suggesting that Vid24p acts after FBPase sequestration into the vesicles has occurred. Vid24p is the first protein identified that is localized to the FBPase-containing vesicles and plays a critical role in delivering FBPase from the vesicles to the vacuole for degradation.  相似文献   

3.
Kallikrein 4 (Klk4) is believed to play an essential role in enamel biomineralization, because defects in KLK4 cause hypomaturation amelogenesis imperfecta. We used gene targeting to generate a knockin mouse that replaces the Klk4 gene sequence, starting at the translation initiation site, with a lacZ reporter gene. Correct targeting of the transgene was confirmed by Southern blot and PCR analyses. Histochemical X-gal (5-bromo-4-chloro-3-indolyl-β-d-galactopyranoside) staining demonstrated expression of β-galactosidase in maturation stage ameloblasts. No X-gal staining was observed in secretory stage ameloblasts or in odontoblasts. Retained enamel proteins were observed in the maturation stage enamel of the Klk4 null mouse, but not in the Klk4 heterozygous or wild-type mice. The enamel layer in the Klk4 null mouse was normal in thickness and contained decussating enamel rods but was rapidly abraded following weaning, despite the mice being maintained on soft chow. In function the enamel readily fractured within the initial rod and interrod enamel above the parallel enamel covering the dentino-enamel junction. Despite the lack of Klk4 and the retention of enamel proteins, significant levels of crystal maturation occurred (although delayed), and the enamel achieved a mineral density in some places greater than that detected in bone and dentin. An important finding was that individual enamel crystallites of erupted teeth failed to grow together, interlock, and function as a unit. Instead, individual crystallites seemed to spill out of the enamel when fractured. These results demonstrate that Klk4 is essential for the removal of enamel proteins and the proper maturation of enamel crystals.Dental enamel is composed of highly ordered, very long crystals of calcium hydroxyapatite (Ca10(PO4)6(OH)2). Mature enamel crystallites are about 70 nm wide and 30 nm thick, but are of unmeasurable length (1), probably extending all the way from the dentin layer to the surface of the tooth (2). Enamel crystallites are organized into bundles called rods, with about 10,000 parallel crystals in a rod (3). Each enamel rod is the product of a single ameloblast, the cell type that forms a continuous sheet over the developing enamel and orchestrates its formation. Dental enamel of erupted teeth is ∼95% mineral (by weight) (4), with most of the non-mineral component being water. Protein comprises <1% of its weight. Forming enamel, however, is over 30% protein (5). Much of the protein is reabsorbed by ameloblasts and degraded in lysosomes (6, 7), but extracellular proteases also play a role in matrix protein removal (810).Dental enamel formation is divided into secretory, transition, and maturation stages (11, 12). During the secretory stage, enamel crystals grow primarily in length. As the crystals extend, the enamel layer expands. Enamel crystallites lengthen along a mineralization front at the secretory surface of the ameloblast cell membrane. There, mineral deposits rapidly on the crystallite tips, and very slowly on their sides (3, 13, 14). By the end of the secretory stage the enamel crystals are full-length and the enamel layer as a whole is as thick as it will ever be, but it has only about 14% of the mineral as it will have when the tooth erupts (15). Following the secretory stage there is then a transition during which the ameloblasts greatly reduce their secretion of enamel proteins (16) and convert to maturation ameloblasts (17). During the maturation stage, mineral is deposited exclusively on the sides of pre-existing enamel crystallites (18), which grow in width and thickness until further growth is prevented by contact with adjacent crystals (19, 20). During early maturation the percentage protein by weight drops from 30 to 2% (5), and half of the total enamel mineral is deposited. The final 30–35% of mineral is deposited in the absence of significant protein and allows the crystals to grow firmly against one another and to mechanically interlock (15).The major secretory stage enamel proteins are amelogenin (21, 22), ameloblastin (2325), and enamelin (26, 27). These proteins function specifically during enamel formation, and the disease phenotypes exhibited by mice lacking these genes are confined to the developing teeth and include enamel agenesis (2830). These genes are often deleted or are reduced to pseudogenes in vertebrates such as birds or baleen whales that evolved alternatives to developing teeth (31, 32). Although the enamel extracellular matrix proteins are critical for growing enamel crystals, they are not part of the final enamel product. Prior to tooth eruption, enamel proteins are digested by proteases and reabsorbed by ameloblasts. Two extracellular matrix proteases are involved in the cleavage of enamel proteins: matrix metalloproteinase 20 (Mmp-20)2 (33) and kallikrein 4 (Klk4) (34).Mmp-20 is secreted along with amelogenin, ameloblastin, and enamelin by secretory stage ameloblasts (3537). Mmp-20 activity can account for the range of cleavages observed in secretory stage enamel proteins (38) and appears to be the only protease secreted by ameloblasts during the secretory stage. Mmp20-null mice have enamel that is thinner and softer than normal, lacks enamel rod organization, and tends to chip off the crown surface (39, 40). Like the other secretory stage enamel proteins, Mmp20 expression appears to be restricted to developing teeth (41), as is the diseased phenotype when the human gene is defective (4244).Klk4 is a serine protease that is secreted by transition and maturation stage ameloblasts but is not expressed by secretory stage ameloblasts (45, 46). Klk4 might also be expressed by odontoblasts, the cells that form dentin (47). Klk4 has broad substrate specificity (48, 49) and is capable of activating other proteases (5052) and protease activated receptors (53, 54). Unlike most proteins secreted by ameloblasts, Klk4 is expressed in other tissues, most notably the prostate (55) and endometrium (56). Much attention has been focused on the potential role of Klk4 in cancers. Klk4 is increased in breast cancer stromal cells (57), in prostate cancer cells (5861), and ovarian cancer cells (6265). Despite this focus on the potential role of Klk4 in tumors, very little is known about the normal expression and function of Klk4 in nondental tissues. A loss of function mutation in both human KLK4 alleles caused a hypomaturation enamel phenotype in the absence of any observable defects elsewhere in the body (66).To gain insights into the role of Klk4 in normal dental enamel formation, and to better characterize the normal temporal and spatial patterns of Klk4 expression, we have used gene targeting to knock out normal Klk4 expression, while replacing the Klk4 code with lacZ, the bacterial gene encoding β-galactosidase reporter in mice. We demonstrate that Klk4 is not expressed by secretory stage ameloblasts, but is specifically expressed by ameloblasts later in enamel formation and is necessary for the proper removal of enamel proteins, the final thickening of enamel crystals, and ultimately, for hardening of the enamel layer.  相似文献   

4.
5.
Leptospira spp., the causative agents of leptospirosis, adhere to components of the extracellular matrix, a pivotal role for colonization of host tissues during infection. Previously, we and others have shown that Leptospira immunoglobulin-like proteins (Lig) of Leptospira spp. bind to fibronectin, laminin, collagen, and fibrinogen. In this study, we report that Leptospira can be immobilized by human tropoelastin (HTE) or elastin from different tissues, including lung, skin, and blood vessels, and that Lig proteins can bind to HTE or elastin. Moreover, both elastin and HTE bind to the same LigB immunoglobulin-like domains, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12 as demonstrated by enzyme-linked immunosorbent assay (ELISA) and competition ELISAs. The LigB immunoglobulin-like domain binds to the 17th to 27th exons of HTE (17–27HTE) as determined by ELISA (LigBCon4, KD = 0.50 μm; LigBCen7′–8, KD = 0.82 μm; LigBCen9, KD = 1.54 μm; and LigBCen12, KD = 0.73 μm). The interaction of LigBCon4 and 17–27HTE was further confirmed by steady state fluorescence spectroscopy (KD = 0.49 μm) and ITC (KD = 0.54 μm). Furthermore, the binding was enthalpy-driven and affected by environmental pH, indicating it is a charge-charge interaction. The binding affinity of LigBCon4D341N to 17–27HTE was 4.6-fold less than that of wild type LigBCon4. In summary, we show that Lig proteins of Leptospira spp. interact with elastin and HTE, and we conclude this interaction may contribute to Leptospira adhesion to host tissues during infection.Pathogenic Leptospira spp. are spirochetes that cause leptospirosis, a serious infectious disease of people and animals (1, 2). Weil syndrome, the severe form of leptospiral infection, leads to multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, meningitis, abortion, and uveitis (3, 4). Furthermore, this disease is not only prevalent in many developing countries, it is reemerging in the United States (3). Although leptospirosis is a serious worldwide zoonotic disease, the pathogenic mechanisms of Leptospira infection remain enigmatic. Recent breakthroughs in applying genetic tools to Leptospira may facilitate studies on the molecular pathogenesis of leptospirosis (58).The attachment of pathogenic Leptospira spp. to host tissues is critical in the early phase of Leptospira infection. Leptospira spp. adhere to host tissues to overcome mechanical defense systems at tissue surfaces and to initiate colonization of specific tissues, such as the lung, kidney, and liver. Leptospira invade hosts tissues through mucous membranes or injured epidermis, coming in contact with subepithelial tissues. Here, certain bacterial outer surface proteins serve as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 to mediate the binding of bacteria to different extracellular matrices (ECMs) of host cells (9). Several leptospiral MSCRAMMs have been identified (1018), and we speculate that more will be identified in the near future.Lig proteins are distributed on the outer surface of pathogenic Leptospira, and the expression of Lig protein is only found in low passage strains (14, 16, 17), probably induced by environmental cues such as osmotic or temperature changes (19). Lig proteins can bind to fibrinogen and a variety of ECMs, including fibronectin (Fn), laminin, and collagen, thereby mediating adhesion to host cells (2023). Lig proteins also constitute good vaccine candidates (2426).Elastin is a component of ECM critical to tissue elasticity and resilience and is abundant in skin, lung, blood vessels, placenta, uterus, and other tissues (2729). Tropoelastin is the soluble precursor of elastin (28). During the major phase of elastogenesis, multiple tropoelastin molecules associate through coacervation (3032). Because of the abundance of elastin or tropoelastin on the surface of host cells, several bacterial MSCRAMMs use elastin and/or tropoelastin to mediate adhesion during the infection process (3335).Because leptospiral infection is known to cause severe pulmonary hemorrhage (36, 37) and abortion (38), we hypothesize that some leptospiral MSCRAMMs may interact with elastin and/or tropoelastin in these elastin-rich tissues. This is the first report that Lig proteins of Leptospira interact with elastin and tropoelastin, and the interactions are mediated by several specific immunoglobulin-like domains of Lig proteins, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12, which bind to the 17th to 27th exons of human tropoelastin (HTE).  相似文献   

6.
7.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

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

10.
11.
12.
Paneth cells are a secretory epithelial lineage that release dense core granules rich in host defense peptides and proteins from the base of small intestinal crypts. Enteric α-defensins, termed cryptdins (Crps) in mice, are highly abundant in Paneth cell secretions and inherently resistant to proteolysis. Accordingly, we tested the hypothesis that enteric α-defensins of Paneth cell origin persist in a functional state in the mouse large bowel lumen. To test this idea, putative Crps purified from mouse distal colonic lumen were characterized biochemically and assayed in vitro for bactericidal peptide activities. The peptides comigrated with cryptdin control peptides in acid-urea-PAGE and SDS-PAGE, providing identification as putative Crps. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry experiments showed that the molecular masses of the putative α-defensins matched those of the six most abundant known Crps, as well as N-terminally truncated forms of each, and that the peptides contain six Cys residues, consistent with identities as α-defensins. N-terminal sequencing definitively revealed peptides with N termini corresponding to full-length, (des-Leu)-truncated, and (des-Leu-Arg)-truncated N termini of Crps 1–4 and 6. Crps from mouse large bowel lumen were bactericidal in the low micromolar range. Thus, Paneth cell α-defensins secreted into the small intestinal lumen persist as intact and functional forms throughout the intestinal tract, suggesting that the peptides may mediate enteric innate immunity in the colonic lumen, far from their upstream point of secretion in small intestinal crypts.Antimicrobial peptides (AMPs)2 are released by epithelial cells onto mucosal surfaces as effectors of innate immunity (15). In mammals, most AMPs derive from two major families, the cathelicidins and defensins (6). The defensins comprise the α-, β-, and θ-defensin subfamilies, which are defined by the presence of six cysteine residues paired in characteristic tridisulfide arrays (7). α-Defensins are highly abundant in two primary cell lineages: phagocytic leukocytes, primarily neutrophils, of myeloid origin and Paneth cells, which are secretory epithelial cells located at the base of the crypts of Lieberkühn in the small intestine (810). Neutrophil α-defensins are stored in azurophilic granules and contribute to non-oxidative microbial cell killing in phagolysosomes (11, 12), except in mice whose neutrophils lack defensins (13). In the small bowel, α-defensins and other host defense proteins (1418) are released apically as components of Paneth cell secretory granules in response to cholinergic stimulation and after exposure to bacterial antigens (19). Therefore, the release of Paneth cell products into the crypt lumen is inferred to protect mitotically active crypt cells from colonization by potential pathogens and confer protection against enteric infection (7, 20, 21).Under normal, homeostatic conditions, Paneth cells are not found outside the small bowel, although they may appear ectopically in response to local inflammation throughout the gastrointestinal tract (22, 23). Paneth cell numbers increase progressively throughout the small intestine, occurring at highest numbers in the distal ileum (24). Mouse Paneth cells express numerous α-defensin isoforms, termed cryptdins (Crps) (25), that have broad spectrum antimicrobial activities (6, 26). Collectively, α-defensins constitute approximately seventy percent of the bactericidal peptide activity in mouse Paneth cell secretions (19), selectively killing bacteria by membrane-disruptive mechanisms (2730). The role of Paneth cell α-defensins in gastrointestinal mucosal immunity is evident from studies of mice transgenic for human enteric α-defensin-5, HD-5, which are immune to infection by orally administered Salmonella enterica sv. typhimurium (S. typhimurium) (31).The biosynthesis of mature, bactericidal α-defensins from their inactive precursors requires activation by lineage-specific proteolytic convertases. In mouse Paneth cells, inactive ∼8.4-kDa Crp precursors are processed intracellularly into microbicidal ∼4-kDa Crps by specific cleavage events mediated by matrix metalloproteinase-7 (MMP-7) (32, 33). MMP-7 null mice exhibit increased susceptibility to systemic S. typhimurium infection and decreased clearance of orally administered non-invasive Escherichia coli (19, 32). Although the α-defensin proregions are sensitive to proteolysis, the mature, disulfide-stabilized peptides resist digestion by their converting enzymes in vitro, whether the convertase is MMP-7 (32), trypsin (34), or neutrophil serine proteinases (35). Because α-defensins resist proteolysis in vitro, we hypothesized that Paneth cell α-defensins resist degradation and remain in a functional state in the large bowel, a complex, hostile environment containing varied proteases of both host and microbial origin.Here, we report on the isolation and characterization of a population of enteric α-defensins from the mouse colonic lumen. Full-length and N-terminally truncated Paneth cell α-defensins were identified and are abundant in the distal large bowel lumen.  相似文献   

13.
14.
15.
16.
17.
18.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

19.
Complex I (EC 1.6.99.3) of the bacterium Escherichia coli is considered to be the minimal form of the type I NADH dehydrogenase, the first enzyme complex in the respiratory chain. Because of its small size and relative simplicity, the E. coli enzyme has become a model used to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex. To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus, which encodes the 14 Nuo subunits comprising E. coli complex I. Here we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.

Complex I (NADH:ubiquinone oxidoreductase; EC 1.6.99.3), a type I NADH dehydrogenase that couples the oxidation of NADH to the generation of a proton motive force, is the first enzyme complex of the respiratory chain (2, 35, 47). The Escherichia coli enzyme, considered to be the minimal form of complex I, consists of 14 subunits instead of the 40 to 50 subunits associated with the homologous eukaryotic mitochondrial enzyme (17, 29, 30, 4850). E. coli also possesses a second NADH dehydrogenase, NDH-II, which does not generate a proton motive force (31). E. coli complex I resembles eukaryotic complex I in many ways (16, 17, 30, 49): it performs the same enzymatic reaction and is sensitive to a number of the same inhibitors, it consists of subunits homologous to those found in all proton-translocating NADH:ubiquinone oxidoreductases studied thus far, it is comprised of a large number of subunits relative to the number that comprise other respiratory enzymes, and it contains flavin mononucleotide and FeS center prosthetic groups. Additionally, it possesses an L-shaped topology (14, 22) like that of its Neurospora crassa homolog (27), and it consists of distinct fragments or subcomplexes. Whereas eukaryotic complex I can be dissected into a peripheral arm and a membrane arm, the E. coli enzyme consists of three subcomplexes referred to as the peripheral, connecting, and membrane fragments (29) (Fig. (Fig.1A).1A). The subunit composition of these three fragments correlates approximately with the organization of the 14 structural genes (nuoA to nuoN) (49) of the nuo (for NADH:ubiquinone oxidoreductase) locus (Fig. (Fig.1B),1B), an organization that is conserved in several other bacteria, including Salmonella typhimurium (3), Paracoccus denitrificans (53), Rhodobacter capsulatus (12), and Thermus thermophilus (54). The 5′ half of the locus contains a promoter (nuoP), previously identified and located upstream of nuoA (8, 49), and the majority of genes that encode subunits homologous to the nucleus-encoded subunits of eukaryotic complex I and to subunits of the Alcaligenes eutrophus NAD-reducing hydrogenase (17, 29, 30, 49). In contrast, the 3′ half contains the majority of the genes that encode subunits homologous to the mitochondrion-encoded subunits of eukaryotic complex I and to subunits of the E. coli formate-hydrogen lyase complex (17, 29, 30, 49). Whereas the nuclear homologs NuoE, NuoF, and NuoG constitute the peripheral fragment (also referred to as the NADH dehydrogenase fragment [NDF]), the nuclear homologs NuoB, NuoC, NuoD, and NuoI constitute the connecting fragment. The mitochondrial homologs NuoA, NuoH, NuoJ, NuoK, NuoL, NuoM, and NuoN constitute the membrane fragment (29). E. coli complex I likely evolved by fusion of preexisting protein assemblies constituting modules for electron transfer and proton translocation (1719, 30). Open in a separate windowFIG. 1Schematic of E. coli complex I and the nuo locus. Adapted with permission of the publisher (17, 29, 30, 49). (A) E. coli complex I is comprised of three distinct fragments: the peripheral (light gray), connecting (white), and membrane (dark gray) fragments (17, 29). The peripheral fragment (NDF) is comprised of the nuclear homologs NuoE, -F, and -G and exhibits NADH dehydrogenase activity that oxidizes NADH to NAD+; the connecting fragment is comprised of the nuclear homologs NuoB, -C, -D, and -I; and the membrane fragment is comprised of the mitochondrial homologs NuoA, -H, and -J to -N and catalyzes ubiquinone (Q) to its reduced form (QH2). FMN, flavin mononucleotide. (B) The E. coli nuo locus encodes the 14 Nuo subunits that constitute complex I. The 5′ half of the locus contains a previously identified promoter (nuoP) and the majority of genes that encode the peripheral and connecting subunits (light gray and white, respectively). The 3′ half of the locus contains the majority of the genes encoding the membrane subunits (dark gray). The 3′ end of nuoG encodes a C-Terminal region (CTR) of the NuoG subunit (hatched).Because of its smaller size and relative simplicity, researchers recently have begun to utilize complex I of E. coli, and that of its close relative S. typhimurium, to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex (3, 8, 46) and to investigate the diverse physiological consequences caused by defects in this enzyme (4, 6, 10, 40, 59). Such defects affect the ability of cells to perform chemotaxis (40), to grow on certain carbon sources (4, 6, 10, 40, 57), to survive stationary phase (59), to perform energy-dependent proteolysis (4), to regulate the expression of at least one gene (32), and to maintain virulence (5).To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus. Here, we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.  相似文献   

20.
The peptidoglycan cortex of endospores of Bacillus species is required for maintenance of spore dehydration and dormancy, and the structure of the cortex may also allow it to function in attainment of spore core dehydration. A significant difference between spore and growing cell peptidoglycan structure is the low degree of peptide cross-linking in cortical peptidoglycan; regulation of the degree of this cross-linking is exerted by d,d-carboxypeptidases. We report here the construction of mutant B. subtilis strains lacking all combinations of two and three of the four apparent d,d-carboxypeptidases encoded within the genome and the analysis of spore phenotypic properties and peptidoglycan structure for these strains. The data indicate that while the dacA and dacC products have no significant role in spore peptidoglycan formation, the dacB and dacF products both function in regulating the degree of cross-linking of spore peptidoglycan. The spore peptidoglycan of a dacB dacF double mutant was very highly cross-linked, and this structural modification resulted in a failure to achieve normal spore core dehydration and a decrease in spore heat resistance. A model for the specific roles of DacB and DacF in spore peptidoglycan synthesis is proposed.Peptidoglycan (PG) is the structural element of the bacterial cell wall which determines cell shape and which resists the turgor pressure within the cell. The bacterial endospores produced by species of Bacillus, Clostridium, and several other bacterial genera are modified cells that are able to survive long periods and extreme conditions in a dormant, relatively dehydrated state. The PG wall within the endospore is required for maintenance of the dehydrated state (10, 11), which is the major determinant of spore heat resistance (2, 17, 22). Spore PG appears to be comprised of two distinct though contiguous layers. The thin inner layer, the germ cell wall, appears to have a structure similar to that of the vegetative wall and serves as the initial cell wall of the germinated spore (1, 20, 21, 31). The thicker outer layer, the spore cortex, has a modified structure which may determine its ability to carry out roles specific to the spore, and is rapidly degraded during spore germination (1, 20, 35, 37). The most dramatic of the cortex structural modifications results in partial cleavage or complete removal of ∼75% of the peptide side chains from the glycan strands. Loss of these peptides limits the cross-linking potential of the PG and results in the formation of only one peptide cross-link per 35 disaccharide units in the spore PG, compared to one peptide cross-link per 2.3 to 2.9 disaccharide units in the vegetative PG (1, 20, 36). This low degree of cross-linking has been predicted to give spore PG a flexibility that allows it to have a role in attainment of spore core dehydration (14, 34) in addition to its clear role in maintenance of dehydration. We are studying the structure and mechanism of synthesis of spore PG in an attempt to discern the roles of this structure and its individual components in determining spore properties.A family of proteins called the penicillin-binding proteins (PBPs) polymerizes PG on the external surface of the cell membrane (reviewed in reference 7). The high-molecular-weight (high-MW) members of this family (generally ≥60 kDa) carry the transglycosylase and transpeptidase activities involved in polymerization and cross-linking of the glycan strands. The low-MW PBPs have commonly been found to possess d,d-carboxypeptidase activity. This activity can remove the terminal d-alanine of the peptide side chains and thereby prevent the side chain from serving as a donor in the formation of a peptide cross-link. Analysis of the B. subtilis genome reveals six low-MW PBP-encoding genes: dacA (33), dacB (4), dacC (19), dacF (38), pbpE (23), and pbpX (accession no. Z99112). The four dac gene products exhibit very high sequence similarity to proven d,d-carboxypeptidases, and this activity has been demonstrated in vitro for the dacA and dacB products, PBP5 (12) and PBP5* (32), respectively. The sequences of the pbpE and pbpX products are more distantly related, and no activity has yet been established or ruled out for them.PBP5 is the major penicillin-binding and d,d-carboxypeptidase activity found in vegetative cells (12). Although dacA expression declines significantly during sporulation, a significant amount of PBP5 remains during the time of spore PG synthesis (29). A dacA-null mutation results in no obvious effects on vegetative growth, sporulation, spore characteristics, or spore germination (3, 33). However, loss of PBP5 does result in a reduction of cleavage of peptide side chains from the tetrapeptide to the tripeptide form in the spore PG (20). PBP5* is expressed only during sporulation and only in the mother cell compartment of the sporangium, under the control of the RNA polymerase ςE subunit (4, 5, 28, 29). A dacB-null mutation leading to loss of this d,d-carboxypeptidase results in a fourfold increase in the effective cross-linking of the spore PG (1, 20, 22). This structural change is accompanied by only slight decreases in spore core dehydration and heat resistance (3, 22). The suspected d,d-carboxypeptidase activities of the products of the dacC and dacF genes have not been demonstrated. The latter two genes are expressed only during the postexponential growth phase: dacC is expressed during early stationary phase under the control of ςH (19) and dacF is expressed only within the forespore under the control of ςF (27, 38). Null mutations effecting either gene result in no obvious phenotype and no change in spore PG structure (19, 38).The multiplicity of these proteins in sporulating cells and the lack of effect of loss of some of them suggested redundancy of function among these proteins, a situation observed previously with PBPs of a high-MW class (25, 30, 39). In order to examine this possibility we have constructed mutants lacking multiple low-MW PBPs and have examined their sporulation efficiency, spore PG structure, spore heat resistance and wet density, and spore germination and outgrowth. The present study demonstrates a role for the dacF gene product in synthesis of spore PG, and we also present a model for the roles of the dacB and dacF gene products in spore PG formation.  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