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.
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   

2.
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   

3.
The Notch receptor is critical for proper development where it orchestrates numerous cell fate decisions. The Fringe family of β1,3-N-acetylglucosaminyltransferases are regulators of this pathway. Fringe enzymes add N-acetylglucosamine to O-linked fucose on the epidermal growth factor repeats of Notch. Here we have analyzed the reaction catalyzed by Lunatic Fringe (Lfng) in detail. A mutagenesis strategy for Lfng was guided by a multiple sequence alignment of Fringe proteins and solutions from docking an epidermal growth factor-like O-fucose acceptor substrate onto a homology model of Lfng. We targeted three main areas as follows: residues that could help resolve where the fucose binds, residues in two conserved loops not observed in the published structure of Manic Fringe, and residues predicted to be involved in UDP-N-acetylglucosamine (UDP-GlcNAc) donor specificity. We utilized a kinetic analysis of mutant enzyme activity toward the small molecule acceptor substrate 4-nitrophenyl-α-l-fucopyranoside to judge their effect on Lfng activity. Our results support the positioning of O-fucose in a specific orientation to the catalytic residue. We also found evidence that one loop closes off the active site coincident with, or subsequent to, substrate binding. We propose a mechanism whereby the ordering of this short loop may alter the conformation of the catalytic aspartate. Finally, we identify several residues near the UDP-GlcNAc-binding site, which are specifically permissive toward UDP-GlcNAc utilization.Defects in Notch signaling have been implicated in numerous human diseases, including multiple sclerosis (1), several forms of cancer (2-4), cerebral autosomal dominant arteriopathy with sub-cortical infarcts and leukoencephalopathy (5), and spondylocostal dysostosis (SCD)3 (6-8). The transmembrane Notch signaling receptor is activated by members of the DSL (Delta, Serrate, Lag2) family of ligands (9, 10). In the endoplasmic reticulum, O-linked fucose glycans are added to the epidermal growth factor-like (EGF) repeats of the Notch extracellular domain by protein O-fucosyltransferase 1 (11-13). These O-fucose monosaccharides can be elongated in the Golgi apparatus by three highly conserved β1,3-N-acetylglucosaminyltransferases of the Fringe family (Lunatic (Lfng), Manic (Mfng), and Radical Fringe (Rfng) in mammals) (14-16). The formation of this GlcNAc-β1,3-Fuc-α1, O-serine/threonine disaccharide is necessary and sufficient for subsequent elongation to a tetrasaccharide (15, 19), although elongation past the disaccharide in Drosophila is not yet clear (20, 21). Elongation of O-fucose by Fringe is known to potentiate Notch signaling from Delta ligands and inhibit signaling from Serrate ligands (22). Delta ligands are termed Delta-like (Delta-like1, -2, and -4) in mammals, and the homologs of Serrate are known as Jagged (Jagged1 and -2) in mammals. The effects of Fringe on Drosophila Notch can be recapitulated in Notch ligand in vitro binding assays using purified components, suggesting that the elongation of O-fucose by Fringe alters the binding of Notch to its ligands (21). Although Fringe also appears to alter Notch-ligand interactions in mammals, the effects of elongation of the glycan past the O-fucose monosaccharide is more complicated and appears to be cell type-, receptor-, and ligand-dependent (for a recent review see Ref. 23).The Fringe enzymes catalyze the transfer of GlcNAc from the donor substrate UDP-α-GlcNAc to the acceptor fucose, forming the GlcNAc-β1,3-Fuc disaccharide (14-16). They belong to the GT-A-fold of inverting glycosyltransferases, which includes N-acetylglucosaminyltransferase I and β1,4-galactosyltransferase I (17, 18). The mechanism is presumed to proceed through the abstraction of a proton from the acceptor substrate by a catalytic base (Asp or Glu) in the active site. This creates a nucleophile that attacks the anomeric carbon of the nucleotide-sugar donor, inverting its configuration from α (on the nucleotide sugar) to β (in the product) (24, 25). The enzyme then releases the acceptor substrate modified with a disaccharide and UDP. The Mfng structure (26) leaves little doubt as to the identity of the catalytic residue, which in all likelihood is aspartate 289 in mouse Lfng (we will use numbering for mouse Lunatic Fringe throughout, unless otherwise stated). The structure of Mfng with UDP-GlcNAc soaked into the crystals (26) showed density only for the UDP portion of the nucleotide-sugar donor and no density for two loops flanking either side of the active site. The presence of flexible loops that become ordered upon substrate binding is a common observation with glycosyltransferases in the GT-A fold family (18, 25). Density for the entire donor was observed in the structure of rabbit N-acetylglucosaminyltransferase I (27). In this case, ordering of a previously disordered loop upon UDP-GlcNAc binding may have contributed to increased stability of the donor. In the case of bovine β1,4-galactosyltransferase I, a section of flexible random coil from the apo-structure was observed to change its conformation to α-helical upon donor substrate binding (28). Both loops in Lfng are highly conserved, and we have mutated a number of residues in each to test the hypothesis that they interact with the substrates. The mutagenesis strategy was also guided by docking of an EGF-O-fucose acceptor substrate into the active site of the Lfng model as well as comparison of the Lfng model with a homology model of the β1,3-glucosyltransferase (β3GlcT) that modifies O-fucose on thrombospondin type 1 repeats (29, 30). The β3GlcT is predicted to be a GT-A fold enzyme related to the Fringe family (17, 18, 29).  相似文献   

4.
5.
6.
Many G protein-coupled receptors (GPCRs) recycle after agonist-induced endocytosis by a sequence-dependent mechanism, which is distinct from default membrane flow and remains poorly understood. Efficient recycling of the β2-adrenergic receptor (β2AR) requires a C-terminal PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (PDZbd), an intact actin cytoskeleton, and is regulated by the endosomal protein Hrs (hepatocyte growth factor-regulated substrate). The PDZbd is thought to link receptors to actin through a series of protein interaction modules present in NHERF/EBP50 (Na+/H+ exchanger 3 regulatory factor/ezrin-binding phosphoprotein of 50 kDa) family and ERM (ezrin/radixin/moesin) family proteins. It is not known, however, if such actin connectivity is sufficient to recapitulate the natural features of sequence-dependent recycling. We addressed this question using a receptor fusion approach based on the sufficiency of the PDZbd to promote recycling when fused to a distinct GPCR, the δ-opioid receptor, which normally recycles inefficiently in HEK293 cells. Modular domains mediating actin connectivity promoted receptor recycling with similarly high efficiency as the PDZbd itself, and recycling promoted by all of the domains was actin-dependent. Regulation of receptor recycling by Hrs, however, was conferred only by the PDZbd and not by downstream interaction modules. These results suggest that actin connectivity is sufficient to mimic the core recycling activity of a GPCR-linked PDZbd but not its cellular regulation.G protein-coupled receptors (GPCRs)2 comprise the largest family of transmembrane signaling receptors expressed in animals and transduce a wide variety of physiological and pharmacological information. While these receptors share a common 7-transmembrane-spanning topology, structural differences between individual GPCR family members confer diverse functional and regulatory properties (1-4). A fundamental mechanism of GPCR regulation involves agonist-induced endocytosis of receptors via clathrin-coated pits (4). Regulated endocytosis can have multiple functional consequences, which are determined in part by the specificity with which internalized receptors traffic via divergent downstream membrane pathways (5-7).Trafficking of internalized GPCRs to lysosomes, a major pathway traversed by the δ-opioid receptor (δOR), contributes to proteolytic down-regulation of receptor number and produces a prolonged attenuation of subsequent cellular responsiveness to agonist (8, 9). Trafficking of internalized GPCRs via a rapid recycling pathway, a major route traversed by the β2-adrenergic receptor (β2AR), restores the complement of functional receptors present on the cell surface and promotes rapid recovery of cellular signaling responsiveness (6, 10, 11). When co-expressed in the same cells, the δOR and β2AR are efficiently sorted between these divergent downstream membrane pathways, highlighting the occurrence of specific molecular sorting of GPCRs after endocytosis (12).Recycling of various integral membrane proteins can occur by default, essentially by bulk membrane flow in the absence of lysosomal sorting determinants (13). There is increasing evidence that various GPCRs, such as the β2AR, require distinct cytoplasmic determinants to recycle efficiently (14). In addition to requiring a cytoplasmic sorting determinant, sequence-dependent recycling of the β2AR differs from default recycling in its dependence on an intact actin cytoskeleton and its regulation by the conserved endosomal sorting protein Hrs (hepatocyte growth factor receptor substrate) (11, 14). Compared with the present knowledge regarding protein complexes that mediate sorting of GPCRs to lysosomes (15, 16), however, relatively little is known about the biochemical basis of sequence-directed recycling or its regulation.The β2AR-derived recycling sequence conforms to a canonical PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (henceforth called PDZbd), and PDZ-mediated protein association(s) with this sequence appear to be primarily responsible for its endocytic sorting activity (17-20). Fusion of this sequence to the cytoplasmic tail of the δOR effectively re-routes endocytic trafficking of engineered receptors from lysosomal to recycling pathways, establishing the sufficiency of the PDZbd to function as a transplantable sorting determinant (18). The β2AR-derived PDZbd binds with relatively high specificity to the NHERF/EBP50 family of PDZ proteins (21, 22). A well-established biochemical function of NHERF/EBP50 family proteins is to associate integral membrane proteins with actin-associated cytoskeletal elements. This is achieved through a series of protein-interaction modules linking NHERF/EBP50 family proteins to ERM (ezrin-radixin-moesin) family proteins and, in turn, to actin filaments (23-26). Such indirect actin connectivity is known to mediate other effects on plasma membrane organization and function (23), however, and NHERF/EBP50 family proteins can bind to additional proteins potentially important for endocytic trafficking of receptors (23, 25). Thus it remains unclear if actin connectivity is itself sufficient to promote sequence-directed recycling of GPCRs and, if so, if such connectivity recapitulates the normal cellular regulation of sequence-dependent recycling. In the present study, we took advantage of the modular nature of protein connectivity proposed to mediate β2AR recycling (24, 26), and extended the opioid receptor fusion strategy used successfully for identifying diverse recycling sequences in GPCRs (27-29), to address these fundamental questions.Here we show that the recycling activity of the β2AR-derived PDZbd can be effectively bypassed by linking receptors to ERM family proteins in the absence of the PDZbd itself. Further, we establish that the protein connectivity network can be further simplified by fusing receptors to an interaction module that binds directly to actin filaments. We found that bypassing the PDZ-mediated interaction using either domain is sufficient to mimic the ability of the PDZbd to promote efficient, actin-dependent recycling of receptors. Hrs-dependent regulation, however, which is characteristic of sequence-dependent recycling of wild-type receptors, was recapitulated only by the fused PDZbd and not by the proposed downstream interaction modules. These results support a relatively simple architecture of protein connectivity that is sufficient to mimic the core recycling activity of the β2AR-derived PDZbd, but not its characteristic cellular regulation. Given that an increasing number of GPCRs have been shown to bind PDZ proteins that typically link directly or indirectly to cytoskeletal elements (17, 27, 30-32), the present results also suggest that actin connectivity may represent a common biochemical principle underlying sequence-dependent recycling of various GPCRs.  相似文献   

7.
Wnt11 signals through both canonical (β-catenin) and non-canonical pathways and is up-regulated during osteoblast differentiation and fracture healing. In these studies, we evaluated the role of Wnt11 during osteoblastogenesis. Wnt11 overexpression in MC3T3E1 pre-osteoblasts increases β-catenin accumulation and promotes bone morphogenetic protein (BMP)-induced expression of alkaline phosphatase and mineralization. Wnt11 dramatically increases expression of the osteoblast-associated genes Dmp1 (dentin matrix protein 1), Phex (phosphate-regulating endopeptidase homolog), and Bsp (bone sialoprotein). Wnt11 also increases expression of Rspo2 (R-spondin 2), a secreted factor known to enhance Wnt signaling. Overexpression of Rspo2 is sufficient for increasing Dmp1, Phex, and Bsp expression and promotes bone morphogenetic protein-induced mineralization. Knockdown of Rspo2 abrogates Wnt11-mediated osteoblast maturation. Antagonism of T-cell factor (Tcf)/β-catenin signaling with dominant negative Tcf blocks Wnt11-mediated expression of Dmp1, Phex, and Rspo2 and decreases mineralization. However, dominant negative Tcf fails to block the osteogenic effects of Rspo2 overexpression. These studies show that Wnt11 signals through β-catenin, activating Rspo2 expression, which is then required for Wnt11-mediated osteoblast maturation.Wnt signaling is a key regulator of osteoblast differentiation and maturation. In mesenchymal stem cell lines, canonical Wnt signaling by Wnt10b enhances osteoblast differentiation (1). Canonical Wnt signaling through β-catenin has also been shown to enhance the chondroinductive and osteoinductive properties of BMP22 (2, 3). During BMP2-induced osteoblast differentiation of mesenchymal stem cell lines, cross-talk between BMP and Wnt pathways converges through the interaction of Smad4 with β-catenin (2).Canonical Wnt signaling is also critical for skeletal development and homeostasis. During limb development, expression of Wnt3a in the apical ectodermal ridge of limb buds maintains cells in a highly proliferative and undifferentiated state (4, 5). Disruption of canonical Wnt signaling in Lrp5/Lrp6 compound knock-out mice results in limb- and digit-patterning defects (6). Wnt signaling is also involved in the maintenance of post-natal bone mass. Gain of function in the Wnt co-receptor Lrp5 leads to increased bone mass, whereas loss of Lrp5 function is associated with decreased bone mass and osteoporosis pseudoglioma syndrome (7, 8). Mice with increased Wnt10b expression have increased trabecular bone, whereas Wnt10b-deficient mice have reduced trabecular bone (9). Similarly, mice nullizygous for the Wnt antagonist sFrp1 have increased trabecular bone accrual throughout adulthood (10).Although canonical Wnt signaling regulates osteoblastogenesis and bone formation, the profile of endogenous Wnts that play a role in osteoblast differentiation and maturation is not well described. During development, Wnt11 is expressed in the perichondrium and in the axial skeleton and sternum (11). Wnt11 expression is increased during glucocorticoid-induced osteogenesis (12), indicating a potential role for Wnt11 in osteoblast differentiation. Interestingly, Wnt11 activates both β-catenin-dependent as well as β-catenin-independent signaling pathways (13). Targeted disruption of Wnt11 results in late embryonic/early post-natal death because of cardiac dysfunction (14). Although these mice have no reported skeletal developmental abnormalities, early lethality obfuscates a detailed examination of post-natal skeletal modeling and remodeling.In murine development, Wnt11 expression overlaps with the expression of R-spondin 2 (Rspo2) in the apical ectodermal ridge (11, 15). R-spondins are a novel family of proteins that share structural features, including two conserved cysteinerich furin-like domains and a thrombospondin type I repeat (16). The four R-spondin family members can activate canonical Wnt signaling (15, 1719). Rspo3 interacts with Frizzled 8 and Lrp6 and enhances Wnt ligand signaling. Rspo1 enhances Wnt signaling by interacting with Lrp6 and inhibiting Dkk-mediated receptor internalization (20). Rspo1 was also shown to potentiate Wnt3a-mediated osteoblast differentiation (21). Rspo2 knock-out mice, which die at birth, have limb patterning defects associated with altered β-catenin signaling (2224). However, the role of Rspo2 in osteoblast differentiation and maturation remains unclear.Herein we report that Wnt11 overexpression in MC3T3E1 pre-osteoblasts activates β-catenin and augments BMP-induced osteoblast maturation and mineralization. Wnt11 increases the expression of Rspo2. Overexpression of Rspo2 in MC3T3E1 is sufficient for augmenting BMP-induced osteoblast maturation and mineralization. Although antagonism of Tcf/β-catenin signaling blocks the osteogenic effects of Wnt11, Rspo2 rescues this block, and knockdown of Rspo2 shows that it is required for Wnt11-mediated osteoblast maturation and mineralization. These studies identify both Wnt11 and Rspo2 as novel mediators of osteoblast maturation and mineralization.  相似文献   

8.
9.
Gluconeogenesis is an important metabolic pathway, which produces glucose from noncarbohydrate precursors such as organic acids, fatty acids, amino acids, or glycerol. Fructose-1,6-bisphosphatase, a key enzyme of gluconeogenesis, is found in all organisms, and five different classes of these enzymes have been identified. Here we demonstrate that Escherichia coli has two class II fructose-1,6-bisphosphatases, GlpX and YggF, which show different catalytic properties. We present the first crystal structure of a class II fructose-1,6-bisphosphatase (GlpX) determined in a free state and in the complex with a substrate (fructose 1,6-bisphosphate) or inhibitor (phosphate). The crystal structure of the ligand-free GlpX revealed a compact, globular shape with two α/β-sandwich domains. The core fold of GlpX is structurally similar to that of Li+-sensitive phosphatases implying that they have a common evolutionary origin and catalytic mechanism. The structure of the GlpX complex with fructose 1,6-bisphosphate revealed that the active site is located between two domains and accommodates several conserved residues coordinating two metal ions and the substrate. The third metal ion is bound to phosphate 6 of the substrate. Inorganic phosphate strongly inhibited activity of both GlpX and YggF, and the crystal structure of the GlpX complex with phosphate demonstrated that the inhibitor molecule binds to the active site. Alanine replacement mutagenesis of GlpX identified 12 conserved residues important for activity and suggested that Thr90 is the primary catalytic residue. Our data provide insight into the molecular mechanisms of the substrate specificity and catalysis of GlpX and other class II fructose-1,6-bisphosphatases.Fructose-1,6-bisphosphatase (FBPase,2 EC 3.1.3.11), a key enzyme of gluconeogenesis, catalyzes the hydrolysis of fructose 1,6-bisphosphate to form fructose 6-phosphate and orthophosphate. It is the reverse of the reaction catalyzed by phosphofructokinase in glycolysis, and the product, fructose 6-phosphate, is an important precursor in various biosynthetic pathways (1). In all organisms, gluconeogenesis is an important metabolic pathway that allows the cells to synthesize glucose from noncarbohydrate precursors, such as organic acids, amino acids, and glycerol. FBPases are members of the large superfamily of lithium-sensitive phosphatases, which includes three families of inositol phosphatases and FBPases (the phosphoesterase clan CL0171, 3167 sequences, Pfam data base). These enzymes show metal-dependent and lithium-sensitive phosphomonoesterase activity and include inositol polyphosphate 1-phosphatases, inositol monophosphatases (IMPases), 3′-phosphoadenosine 5′-phosphatases (PAPases), and enzymes acting on both inositol 1,4-bisphosphate and PAP (PIPases) (2). They possess a common structural core with the active site lying between α+β and α/β domains (3). Li+-sensitive phosphatases are putative targets for lithium therapy in the treatment of manic depressive patients (4), whereas FBPases are targets for the development of drugs for the treatment of noninsulin-dependent diabetes (5, 6). In addition, FBPase is required for virulence in Mycobacterium tuberculosis and Leishmania major and plays an important role in the production of lysine and glutamate by Corynebacterium glutamicum (7, 8).Presently, five different classes of FBPases have been proposed based on their amino acid sequences (FBPases I to V) (911). Eukaryotes contain only the FBPase I-type enzyme, but all five types exist in various prokaryotes. Types I, II, and III are primarily in bacteria, type IV in archaea (a bifunctional FBPase/inositol monophosphatase), and type V in thermophilic prokaryotes from both domains (11). Many organisms have more than one FBPase, mostly the combination of types I + II or II + III, but no bacterial genome has a combination of types I and III FBPases (9). The type I FBPase is the most widely distributed among living organisms and is the primary FBPase in Escherichia coli, most bacteria, a few archaea, and all eukaryotes (9, 1115). The type II FBPases are represented by the E. coli GlpX and FBPase F-I from Synechocystis PCC6803 (9, 16); type III is represented by the Bacillus subtilis FBPase (17); type IV is represented by the dual activity FBPases/inosine monophosphatases FbpA from Pyrococcus furiosus (18), MJ0109 from Methanococcus jannaschii (19), and AF2372 from Archaeoglobus fulgidus (20); and type V is represented by the FBPases TK2164 from Pyrococcus (Thermococcus) kodakaraensis and ST0318 from Sulfolobus tokodai (10, 21).Three-dimensional structures of the type I (from pig kidney, spinach chloroplasts, and E. coli), type IV (MJ0109 and AF2372), and type V (ST0318) FBPases have been solved (10, 11, 19, 20, 22, 23). FBPases I and IV and inositol monophosphatases share a common sugar phosphatase fold organized in five layered interleaved α-helices and β-sheets (α-β-α-β-α) (2, 19, 24). ST0318 (an FBPase V enzyme) is composed of one domain with a completely different four-layer α-β-β-α fold (10). The FBPases from these three classes (I, IV, and V) require divalent cations for activity (Mg2+, Mn2+, or Zn2+), and their structures have revealed the presence of three or four metal ions in the active site.E. coli has five Li+-sensitive phosphatases as follows: CysQ (a PAPase), SuhB (an IMPase), Fbp (a FBPase I enzyme), GlpX (a FBPase II), and YggF (an uncharacterized protein) (see the Pfam data base). CysQ is a 3′-phosphoadenosine 5′-phosphatase involved in the cysteine biosynthesis pathway (25, 26), whereas SuhB is an inositol monophosphatase (IMPase) that is also known as a suppressor of temperature-sensitive growth phenotypes in E. coli (27, 28). Fbp is required for growth on gluconeogenic substrates and probably represents the main gluconeogenic FBPase (12). This enzyme has been characterized both biochemically and structurally and shown to be inhibited by low concentrations of AMP (IC50 15 μm) (11, 29, 30). The E. coli GlpX, a class II enzyme FBPase, has been shown to possess a Mn2+-dependent FBPase activity (9). The increased expression of glpX from a multicopy plasmid complemented the Fbp- phenotype; however, the glpX knock-out strain grew normally on gluconeogenic substrates (succinate or glycerol) (9).In this study, we present the first structure of a class II FBPase, the E. coli GlpX, in a free state and in the complex with FBP + metals or phosphate. We have demonstrated that the fold of GlpX is similar to that of the lithium-sensitive phosphatases. We have identified the GlpX residues important for activity and proposed a catalytic mechanism. We have also showed that YggF is a third FBPase in E. coli, which has distinct catalytic properties and is more sensitive than GlpX to the inhibition by lithium or phosphate.  相似文献   

10.
Glycodelin is a human glycoprotein with four reported glycoforms, namely glycodelin-A (GdA), glycodelin-F (GdF), glycodelin-C (GdC), and glycodelin-S (GdS). These glycoforms have the same protein core and appear to differ in their N-glycosylation. The glycosylation of GdA is completely different from that of GdS. GdA inhibits proliferation and induces cell death of T cells. However, the glycosylation and immunomodulating activities of GdF and GdC are not known. This study aimed to use ultra-high sensitivity mass spectrometry to compare the glycomes of GdA, GdC, and GdF and to study the relationship between the immunological activity and glycosylation pattern among glycodelin glycoforms. Using MALDI-TOF strategies, the glycoforms were shown to contain an enormous diversity of bi-, tri-, and tetra-antennary complex-type glycans carrying Galβ1–4GlcNAc (lacNAc) and/or GalNAcβ1–4GlcNAc (lacdiNAc) antennae backbones with varying levels of fucose and sialic acid substitution. Interestingly, they all carried a family of Sda (NeuAcα2–3(GalNAcβ1–4)Gal)-containing glycans, which were not identified in the earlier study because of less sensitive methodologies used. Among the three glycodelins, GdA is the most heavily sialylated. Virtually all the sialic acid on GdC is located on the Sda antennae. With the exception of the Sda epitope, the GdC N-glycome appears to be the asialylated counterpart of the GdA/GdF glycomes. Sialidase activity, which may be responsible for transforming GdA/GdF to GdC, was detected in cumulus cells. Both GdA and GdF inhibited the proliferation, induced cell death, and suppressed interleukin-2 secretion of Jurkat cells and peripheral blood mononuclear cells. In contrast, no immunosuppressive effect was observed for GdS and GdC.Glycodelin is a member of the lipocalin family. It consists of 180 amino acid residues (1) with two sites of N-linked glycosylation. There are four reported glycodelin isoforms, namely glycodelin-A (amniotic fluid isoform, GdA),4 glycodelin-F (follicular fluid, GdF), glycodelin-C (cumulus matrix, GdC) and glycodelin-S (seminal plasma, GdS) (25). Among the four glycodelin isoforms, only the N-glycan structures of GdA and GdS have been previously determined. This was achieved using fast atom bombardment mass spectrometry (6, 7). The glycan structures of GdA and GdS are completely different. In GdA, the Asn-28 site carries high mannose, hybrid, and complex-type structures, whereas the second Asn-63 site is exclusively occupied by complex-type glycans (6). The major non-reducing epitopes characterized in the complex-type glycans are Galβ1–4GlcNAc (lacNAc), GalNAcβ1–4GlcNAc (lacdiNAc), NeuAcα2–6Galβ1–4GlcNAc (sialylated lacNAc), NeuAcα2–6GalNAcβ1–4GlcNAc (sialylated lacdiNAc), Galβ1–4(Fucα1–3)GlcNAc (Lewis-x), and GalNAcβ1–4(Fucα1–3)GlcNAc (lacdiNAc analog of the blood group substance Lewis-x) (6). Many of these oligosaccharides are rare in other human glycoproteins. GdS glycans are unusually fucose-rich, and the major complex type glycan structures are bi-antennary glycans with Lewis-x and Lewis-y antennae. Glycosylation of GdS is highly site-specific. Asn-28 contains only high mannose structures, whereas Asn-63 contains only complex type glycans. More than 80% of the complex glycans have 3–5 fucose residues/glycan, and none of the glycans is sialylated, which is unusual for a secreted human glycoprotein (7). The glycan structures of GdF and GdC are not known, although they differ in lectin-binding properties and isoelectric point from the other two glycodelin isoforms (5).Glycans are involved in various intracellular, intercellular, and cell-matrix recognition events (8, 9). Glycosylation determines the biological activities of the glycodelin isoforms (2, 10). For example, both GdA and GdF inhibit the spermatozoa-zona pellucida binding (11) via fucosyltransferase-5 (12), but only the latter inhibits progesterone-induced acrosome reaction, thus preventing a premature acrosome reaction of the spermatozoa. There is evidence that cumulus cells can convert exogenous GdA and -F to GdC, the physicochemical properties of which suggest that it is differently glycosylated compared with GdA/F (5). Moreover, GdC stimulated spermatozoa-zona pellucida binding in a dose-dependent manner, and it effectively displaced sperm-bound GdA and -F (4, 5). GdS suppresses capacitation probably via its inhibitory activity on cholesterol efflux from spermatozoa (13).Except for the effects on fertilization, GdA is involved in fetomaternal defense. This glycodelin isoform suppresses proliferation and induces apoptosis of T cells (2) and inhibits natural killer cell (14) and B-cell (15) activities. Glycosylation is involved in the binding of GdA to receptors on T cells (16). The sialic acid of GdA contributes to the apoptotic activity in T cells (17, 18) and binding to CD45, a potential GdA receptor (16). The importance of glycosylation in glycodelin is further shown by the absence of immunosuppressive activities in GdS with different glycosylation (18). The immunomodulating activities of GdF and GdC are unknown.Our previous work showed that glycans are indispensable for the different glycodelins to exhibit their binding activities and biological effects (13, 19, 20). The present study aims to identify the effect of all four glycodelin isoforms on lymphocyte viability, cell death, and interleukin-2 (IL-2) secretion and to correlate these bioactivities with their glycosylation patterns determined by mass spectrometry.  相似文献   

11.
12.
13.
Melanoma cells express the chemokine receptor CXCR4 that confers high invasiveness upon binding to its ligand CXCL12. Melanoma cells at initial stages of the disease show reduction or loss of E-cadherin expression, but recovery of its expression is frequently found at advanced phases. We overexpressed E-cadherin in the highly invasive BRO lung metastatic cell melanoma cell line to investigate whether it could influence CXCL12-promoted cell invasion. Overexpression of E-cadherin led to defective invasion of melanoma cells across Matrigel and type I collagen in response to CXCL12. A decrease in individual cell migration directionality toward the chemokine and reduced adhesion accounted for the impaired invasion. A p190RhoGAP-dependent inhibition of RhoA activation was responsible for the impairment in chemokine-stimulated E-cadherin melanoma transfectant invasion. Furthermore, we show that p190RhoGAP and p120ctn associated predominantly on the plasma membrane of cells overexpressing E-cadherin, and that E-cadherin-bound p120ctn contributed to RhoA inactivation by favoring p190RhoGAP-RhoA association. These results suggest that melanoma cells at advanced stages of the disease could have reduced metastatic potency in response to chemotactic stimuli compared with cells lacking E-cadherin, and the results indicate that p190RhoGAP is a central molecule controlling melanoma cell invasion.Cadherins are a family of Ca2+-dependent adhesion molecules that mediate cell-cell contacts and are expressed in most solid tissues providing a tight control of morphogenesis (1, 2). Classical cadherins, such as epithelial (E) cadherin, are found in adherens junctions, forming core protein complexes with β-catenin, α-catenin, and p120 catenin (p120ctn). Both β-catenin and p120ctn directly interact with E-cadherin, whereas α-catenin associates with the complex through its binding to β-catenin, providing a link with the actin cytoskeleton (1, 2). E-cadherin is frequently lost or down-regulated in many human tumors, coincident with morphological epithelial to mesenchymal transition and acquisition of invasiveness (3-6).Although melanoma only accounts for 5% of skin cancers, when metastasis starts, it is responsible for 80% of deaths from skin cancers (7). Melanocytes express E-cadherin (8-10), but melanoma cells at early radial growth phase show a large reduction in the expression of this cadherin, and surprisingly, expression has been reported to be partially recovered by vertical growth phase and metastatic melanoma cells (9, 11, 12).Trafficking of cancer cells from primary tumor sites to intravasation into blood circulation and later to extravasation to colonize distant organs requires tightly regulated directional cues and cell migration and invasion that are mediated by chemokines, growth factors, and adhesion molecules (13). Solid tumor cells express chemokine receptors that provide guidance of these cells to organs where their chemokine ligands are expressed, constituting a homing model resembling the one used by immune cells to exert their immune surveillance functions (14). Most solid cancer cells express CXCR4, a receptor for the chemokine CXCL12 (also called SDF-1), which is expressed in lungs, bone marrow, and liver (15). Expression of CXCR4 in human melanoma has been detected in the vertical growth phase and on regional lymph nodes, which correlated with poor prognosis and increased mortality (16, 17). Previous in vivo experiments have provided evidence supporting a crucial role for CXCR4 in the metastasis of melanoma cells (18).Rho GTPases control the dynamics of the actin cytoskeleton during cell migration (19, 20). The activity of Rho GTPases is tightly regulated by guanine-nucleotide exchange factors (GEFs),4 which stimulate exchange of bound GDP by GTP, and inhibited by GTPase-activating proteins (GAPs), which promote GTP hydrolysis (21, 22), whereas guanine nucleotide dissociation inhibitors (GDIs) appear to mediate blocking of spontaneous activation (23). Therefore, cell migration is finely regulated by the balance between GEF, GAP, and GDI activities on Rho GTPases. Involvement of Rho GTPases in cancer is well documented (reviewed in Ref. 24), providing control of both cell migration and growth. RhoA and RhoC are highly expressed in colon, breast, and lung carcinoma (25, 26), whereas overexpression of RhoC in melanoma leads to enhancement of cell metastasis (27). CXCL12 activates both RhoA and Rac1 in melanoma cells, and both GTPases play key roles during invasion toward this chemokine (28, 29).Given the importance of the CXCL12-CXCR4 axis in melanoma cell invasion and metastasis, in this study we have addressed the question of whether changes in E-cadherin expression on melanoma cells might affect cell invasiveness. We show here that overexpression of E-cadherin leads to impaired melanoma cell invasion to CXCL12, and we provide mechanistic characterization accounting for the decrease in invasion.  相似文献   

14.
The intestinal epithelium is repetitively deformed by shear, peristalsis, and villous motility. Such repetitive deformation stimulates the proliferation of intestinal epithelial cells on collagen or laminin substrates via ERK, but the upstream mediators of this effect are poorly understood. We hypothesized that the phosphatidylinositol 3-kinase (PI3K)/AKT cascade mediates this mitogenic effect. PI3K, AKT, and glycogen synthase kinase-3β (GSK-3β) were phosphorylated by 10 cycles/min strain at an average 10% deformation, and pharmacologic blockade of these molecules or reduction by small interfering RNA (siRNA) prevented the mitogenic effect of strain in Caco-2 or IEC-6 intestinal epithelial cells. Strain MAPK activation required PI3K but not AKT. AKT isoform-specific siRNA transfection demonstrated that AKT2 but not AKT1 is required for GSK-3β phosphorylation and the strain mitogenic effect. Furthermore, overexpression of AKT1 or an AKT chimera including the PH domain and hinge region of AKT2 and the catalytic domain and C-tail of AKT1 prevented strain activation of GSK-3β, but overexpression of AKT2 or a chimera including the PH domain and hinge region of AKT1 and the catalytic domain and C-tail of AKT2 did not. These data delineate a role for PI3K, AKT2, and GSK-3β in the mitogenic effect of strain. PI3K is required for both ERK and AKT2 activation, whereas AKT2 is sequentially required for GSK-3β. Furthermore, AKT2 specificity requires its catalytic domain and tail region. Manipulating this pathway may prevent mucosal atrophy and maintain the mucosal barrier in conditions such as ileus, sepsis, and prolonged fasting when peristalsis and villous motility are decreased and the mucosal barrier fails.Mechanical forces are part of the normal intestinal epithelial environment. Numerous different forces deform these cells including shear stress from endoluminal chyme, bowel peristalsis, and villous motility (1, 2). During normal bowel function the mucosa is subjected to injury that must be repaired to maintain the mucosal barrier (3, 4). Deformation patterns of the bowel are altered in conditions such as prolonged fasting, post-surgical ileus, and sepsis states, resulting in profoundly reduced mucosal deformation. When such states are prolonged, proliferation slows, the mucosa becomes atrophic, and bacterial translocation may ensue as the mucosal barrier of the gut breaks down (57).In vitro, repetitive deformation is trophic for intestinal epithelial cells (8) cultured on type I or type IV collagen or laminin. Human Caco-2 intestinal epithelial cells (9), non-transformed rat IEC-6 intestinal epithelial cells (10), and primary human intestinal epithelial cells isolated from surgical specimens (11) proliferate more rapidly in response to cyclic strain (12) unless substantial quantities of fibronectin are added to the media or matrix (11) to mimic the acute phase reaction of acute or chronic inflammation and injury. Cyclic strain also stimulates proliferation in HCT 116 colon cancer cells (13) and differentiation of Caco-2 cells cultured on a collagen substrate (9). This phenomenon has also been observed in vivo (14). Thus, repetitive deformation may help to maintain the normal homeostasis of the gut mucosa under non-inflammatory conditions. Previous work in our laboratory has implicated Src, focal adhesion kinase, and the mitogen-activated protein kinase (MAPK)2 extracellular signal-related kinase (ERK) in the mitogenic effect of strain (10). Although p38 is also activated in Caco-2 cells subjected to cyclic strain on a collagen matrix, its activity is not required for the mitogenic effect of strain (12).Although often the PI3K/AKT pathway is thought of as a parallel pathway to the MAPK, this is not always the case. Protein kinase C isoenzymes differentially modulate thrombin effect on MAPK-dependent retinal pigment epithelial cell (RPE) proliferation, and it has been shown that PI3K or AKT inhibition prevented thrombin-induced ERK activation and RPE proliferation (15).PI3K, AKT, and glycogen synthase kinase (GSK), a downstream target of AKT (16), have been implemented in intestinal epithelial cell proliferation in numerous cell systems not involving strain (1719) including uncontrolled proliferation in gastrointestinal cancers (2022). Mechanical forces activate this pathway as well. PI3K and AKT are required for increased extracellular pressure to stimulate colon cancer cell adhesion (23), although the pathway by which pressure stimulates colon cancer cells in suspension differs from the response of adherent intestinal epithelial cells to repetitive deformation (24), and GSK is not involved in this effect.3 Repetitive strain also stimulates vascular endothelial cell proliferation via PI3K and AKT (25, 26), whereas respiratory strain stimulates angiogenic responses via PI3K (27). We, therefore, hypothesized that the PI3K/AKT/GSK axis would be involved in the mitogenic effects of repetitive deformation on a collagen matrix.To test this hypothesis, we used the Flexcell apparatus to rhythmically deform Caco-2 intestinal epithelial cells. IEC-6 cells were used to confirm key results. A frequency of 10 cycles per min was used, which is similar in order of magnitude to the frequency that the intestinal mucosa might be deformed by peristalsis or villous motility in vivo (28, 29). Mechanical forces such as repetitive deformation are likely cell-type and frequency-specific, as different cell types respond to different frequencies. Vascular endothelial cells respond to frequencies of 60–80 cycles/min (25), whereas intestinal epithelial cells may actually decrease proliferation in response to frequencies of 5 cycles/min (30). We characterized PI3K, AKT, and GSK phosphorylation with strain, blocked these molecules pharmacologically or by siRNA, and delineated the specificity of the AKT effect using isozyme-specific siRNA and transfection of AKT1/2 chimeras. We also characterized the interaction of this pathway with the activation of ERK by strain, which has previously been implicated in the mitogenic response (12).  相似文献   

15.
Filaggrin is a component of the cornified cell envelope and the precursor of free amino acids acting as a natural moisturizing factor in the stratum corneum. Deimination is critical for the degradation of filaggrin into free amino acids. In this study, we tried to identify the enzyme(s) responsible for the cleavage of deiminated filaggrin in vitro. First, we investigated citrulline aminopeptidase activity in the extract of newborn rat epidermis by double layer fluorescent zymography and detected strong activity at neutral pH. Monitoring the citrulline-releasing activity, we purified an enzyme of 280 kDa, comprised of six identical subunits of 48 kDa. The NH2 terminus of representative tryptic peptides perfectly matched the sequence of rat bleomycin hydrolase (BH). The enzyme released various amino acids except Pro from β-naphthylamide derivatives and hydrolyzed citrulline-β-naphthylamide most effectively. Thus, to break down deiminated filaggrin, another protease would be required. Among proteases tested, calpain I degraded the deiminated filaggrin effectively into many peptides of different mass on the matrix-assisted laser desorption/ionization-time of flight mass spectrum. We confirmed that various amino acids including citrulline were released by BH from those peptides. On the other hand, caspase 14 degraded deiminated filaggrin into a few peptides of limited mass. Immunohistochemical analysis of normal human skin revealed co-localization of BH and filaggrin in the granular layer. Collectively, our results suggest that BH is essential for the synthesis of natural moisturizing factors and that calpain I would play a role as an upstream protease in the degradation of filaggrin.The mammalian epidermal keratinocytes arise from proliferating basal cells and move outward through a series of distinct differentiation events to form the stratum corneum (1, 2). During this progressive epidermal differentiation, keratinocytes express different proteins such as keratins, profilaggrin/filaggrin, involucrin, small proline-rich proteins, loricrin, cystatin A, and elafin, which form the cornified envelope of mature corneocytes (37). Profilaggrin is synthesized as a large, extremely insoluble phosphoprotein that consists of a unique NH2-terminal Ca2+-binding protein of the S-100 family, linked to 10–20 tandem filaggrin monomer repeats (810). Each individual filaggrin repeat is completely removed by proteolysis to generate the mature filaggrin monomer (a molecular mass of 37 kDa in human). Then, filaggrin is completely degraded in the uppermost layer of the stratum corneum to produce a mixture of free and modified hygroscopic amino acids that are important for maintaining epidermal hydration (2, 1113). In addition, a number of proteins are subjected to various post-translational modifications such as disulfide bonding, N-(γ-glutamyl)-lysine isopeptide cross-linking, and deimination during the terminal differentiation of epidermal keratinocytes (4, 6, 14, 15). Deimination is catalyzed by peptidylarginine deiminase (PAD),2 which converts arginine to citrulline in proteins (1719). The modification seems essential for the processing into free amino acids including citrulline.Several proteases reportedly participate in the processing of profilaggrin. Furin, a member of the proprotein convertase family, has been proposed to cleave the NH2 terminus of profilaggrin, facilitating the release of the NH2-terminal S-100 protein (20, 21). In contrast, calpain I and profilaggrin endopeptidase I (PEP-I) were implicated in the processing of the linker regions between the filaggrin monomer repeats to generate the filaggrin monomer (2225). Recently, significant results regarding the conversion of profilaggrin to filaggrin have been obtained with the knock-out of matriptase/MT-SP1, prostasin/channel-activating serine protease 1/Prss 8, and caspase 14 in mice (2628). These proteases were a key component of the profilaggrin-processing pathway in terminal epidermal differentiation. However, although the signal initiating the degradation of profilaggrin at a defined stage of the maturation of the stratum corneum was found to be the water gradient within the stratum corneum itself (11), the proteases for the processing of filaggrin and/or the deiminated form into peptides following the breakdown of these peptides to amino acids including citrulline remain unknown.In this study, we have purified a novel aminopeptidase using a deiminated substrate from rat skin homogenate and identified it as a neutral cysteine protease, bleomycin hydrolase (BH). Furthermore, we investigated the processing of the deiminated filaggrin by calpain I or caspase 14. Based on these results, we proposed that calpain I participated preferentially in the processing of deiminated filaggrin into peptides and then BH appeared essential for the breakdown of the peptides into amino acids.  相似文献   

16.
17.
Recent studies have ascribed many non-pumping functions to the Na/K-ATPase. We show here that graded knockdown of cellular Na/K-ATPase α1 subunit produces a parallel decrease in both caveolin-1 and cholesterol in light fractions of LLC-PK1 cell lysates. This observation is further substantiated by imaging analyses, showing redistribution of cholesterol from the plasma membrane to intracellular compartments in the knockdown cells. Moreover, this regulation is confirmed in α1+/– mouse liver. Functionally, the knockdown-induced redistribution appears to affect the cholesterol sensing in the endoplasmic reticulum, because it activates the sterol regulatory element-binding protein pathway and increases expression of hydroxymethylglutaryl-CoA reductase and low density lipoprotein receptor in the liver. Consistently, we detect a modest increase in hepatic cholesterol as well as a reduction in the plasma cholesterol. Mechanistically, α1+/– livers show increases in cellular Src and ERK activity and redistribution of caveolin-1. Although activation of Src is not required in Na/K-ATPase-mediated regulation of cholesterol distribution, the interaction between the Na/K-ATPase and caveolin-1 is important for this regulation. Taken together, our new findings demonstrate a novel function of the Na/K-ATPase in control of the plasma membrane cholesterol distribution. Moreover, the data also suggest that the plasma membrane Na/K-ATPase-caveolin-1 interaction may represent an important sensing mechanism by which the cells regulate the sterol regulatory element-binding protein pathway.The Na/K-ATPase, also called the sodium pump, is an ion transporter that mediates active transport of Na+ and K+ across the plasma membrane by hydrolyzing ATP (1, 2). The functional sodium pump is mainly composed of α and β subunits. The α subunit is the catalytic component of the holoenzyme; it contains both the nucleotide and the cation binding sites (3). So far, four isoforms of α subunit have been discovered, and each one shows a distinct tissue distribution pattern (4, 5). Interestingly, studies during the past few years have uncovered many non-pumping functions of Na/K-ATPase (610). Recently, we have demonstrated that more than half of the Na/K-ATPase may actually perform cellular functions other than ion pumping at least in LLC-PK1 cells (11). Moreover, the non-pumping pool of Na/K-ATPase mainly resides in caveolae and interacts with a variety of proteins such as Src, inositol 1,4,5-trisphosphate receptor, and caveolin-1 (1214). While the interaction between Na/K-ATPase and inositol 1,4,5-trisphosphate receptor facilitates Ca2+ signaling (13) the dynamic association between Na/K-ATPase and Src appears to be an essential step for ouabain to stimulate cellular kinases (15). More recently, we report that the interaction between the Na/K-ATPase and caveolin-1 plays an important role for the membrane trafficking of caveolin-1. Knockdown of the Na/K-ATPase leads to altered subcellular distribution of caveolin-1 and increases the mobility of caveolin-1-containing vesicles (16).Caveolin is a protein marker for caveolae (17). Caveolae are flask-shaped vesicular invaginations of plasma membrane and are enriched in cholesterol, glycosphingolipids, and sphingomyelin (18). There are three genes and six isoforms of caveolin. Caveolin-1 is a 22-kDa protein and is expressed in many types of cells, including epithelial and endothelial cells. In addition to their role in biogenesis of caveolae (19), accumulating evidence has implicated caveolin proteins in cellular cholesterol homeostasis (20). For instance, caveolin-1 directly binds to cholesterol in a 1:1 ratio (21). It was also found to be an integral member of the intracellular cholesterol trafficking machinery between internal membranes and plasma membrane (22, 23). The expression of caveolin-1 appears to be under control of SREBPs,2 the master regulators of intracellular cholesterol level (24). Furthermore, knockout of caveolin-1 significantly affected cholesterol metabolism in mouse embryonic fibroblasts and mouse peritoneal macrophages (25). Because we found that the Na/K-ATPase regulates cellular distribution of caveolin-1, we propose that it may also affect intracellular cholesterol distribution and metabolism. To test our hypothesis, we have investigated whether sodium pump α1 knockdown affects cholesterol distribution and metabolism both in vitro and in vivo. Our results indicate that sodium pump α1 expression level plays a role in the proper distribution of intracellular cholesterol. Down-regulation of sodium pump α1 not only redistributes cholesterol between the plasma membrane and cytosolic compartments, but also alters cholesterol metabolism in mice.  相似文献   

18.
The human pathogen Streptococcus pyogenes secretes a highly cytolytic toxin known as streptolysin S (SLS). SLS is a key virulence determinant and responsible for the β-hemolytic phenotype of these bacteria. Despite over a century of research, the chemical structure of SLS remains unknown. Recent experiments have revealed that SLS is generated from an inactive precursor peptide that undergoes extensive post-translational modification to an active form. In this work, we address outstanding questions regarding the SLS biosynthetic process, elucidating the features of substrate recognition and sites of posttranslational modification to the SLS precursor peptide. Further, we exploit these findings to guide the design of artificial cytolytic toxins that are recognized by the SLS biosynthetic enzymes and others that are intrinsically cytolytic. This new structural information has ramifications for future antimicrobial therapies.Streptolysin S (SLS)4 is secreted by the human pathogen Streptococcus pyogenes, the causative agent of diseases ranging from pharyngitis to necrotizing fasciitis (1). SLS is a potent cytolysin that is ribosomally synthesized, extensively posttranslationally modified, and exported to exert its effects on the target cell (2, 3). The expression of SLS promotes virulence in animal models of invasive infection and accounts for the hall-mark zone of β-hemolysis surrounding colonies of these bacteria grown on blood agar (2, 4). An intriguing feature of SLS is its nonimmunogenic nature (5). This characteristic is likely due to its small size and its capacity to lyse cells involved in both innate and adaptive immunity (6, 7). The β-hemolytic phenotype of S. pyogenes has been studied since the early 1900s, but the molecular structure of SLS has remained elusive (8). In the last decade, transposon mutagenesis studies identified the gene encoding the SLS toxin precursor (sagA, for SLS-associated gene) and eight additional genes in an operon required for toxin maturation and export (9). Targeted mutagenesis of the sag operon yields nonhemolytic S. pyogenes mutants with markedly diminished virulence in mice (2). More recently, it was demonstrated that the protein products of sagA–D are sufficient for the in vitro reconstitution of cytolytic activity (3). The first gene product, SagA, serves as a structural template that after a series of tailoring reactions matures into the active SLS metabolite (see Fig. 1A). A trimeric complex of SagBCD catalyzes these tailoring reactions, which results in the conversion of cysteine, serine, and threonine residues to thiazole, oxazole, and methyloxazole heterocycles, respectively (3).Open in a separate windowFIGURE 1.SagBCD substrate recognition is provided by the SagA leader peptide. A, SagA is converted into an active cytolysin, pro-streptolysin-S (pro-SLS), by the actions of SagBCD (a trimeric oxazole/thiazole synthetase). Heterocycles are schematically represented as shaded pentagons. A marginally conserved motif in the SagA leader peptide, FXXXB (where B is a branched chain amino acid), is highlighted in red. Individual reactions catalyzed by SagC (cyclodehydratase) and SagB (FMN-dehydrogenase) are shown. B, representative amino acid sequences and cytolytic activity of SagA-like substrates. Shown in red are leader peptide residues that comprise the FXXXB motif. The putative leader peptide cleavage sites are shown as asterisks, except for McbA, where the site is known (hyphen). In blue are sites of potential heterocycle formation (for McbA, known sites are blue). The percentage of amino acid similarity to full-length SagA (as determined by ClustalW alignment) is given. The cytolytic activity was tested for these substrates in vitro using purified proteins and in vivo using the SLS-deficient strain, S. pyogenes ΔsagA, complemented with the desired substrate. Activity equal to wild type SagA is designated as (+++); activity that is 30–70% of wild type SagA is (++); detectable activity that is less than 30% of SagA is noted as (+); and nondetectable activity is (-). The activity for McbA is not applicable (n.a.) because this secondary metabolite is a DNA gyrase inhibitor, not a cytolysin. C, sequences and lytic activity of mutant substrates. All of the substrates contain the wild type SagA leader peptide, except for the first entry (FXXXB mutant, SagA-FIA). The percentage of amino acid similarity to the protoxin half of SagA is shown. The second and third entries are SagA leader peptides fused to the protoxin of StaphA and ListA. SagX is an artificially designed toxin, whereas the inverse and scrambled substrates manipulate the sequence of SagA between residues 33–50 (underlined).A DNA gyrase inhibitor, microcin B17, is produced by an orthologous biosynthetic cluster (mcb) found in a subset of Escherichia coli strains (1012). Microcin B17 contains four thiazole and four oxazole heterocycles, which are indispensable for biological activity. By analogy to microcin B17 and the lantibiotics, the heterocycles of SLS are formed on the C terminus of SagA, whereas the N terminus serves as a leader peptide (1315). The installation of thiazole and (methyl)-oxazole heterocycles restricts backbone conformational flexibility and provides microcin B17 and SLS with rigidified structures. The SLS heterocycles are formed via two distinct steps; SagC, a cyclodehydratase, generates thiazoline and (methyl)-oxazoline heterocycles, whereas SagB, a dehydrogenase, removes two electrons to afford the aromatic thiazole and (methyl)-oxazole (3, 16, 17). SagD is proposed to play a role in trimer formation and regulation (see Fig. 1A). The final genes in the genetic cluster encode a predicted leader peptidase/immunity protein (SagE), a membrane-associated protein of unknown function (SagF), and three ABC transporters (SagGHI).It is now appreciated that many other prokaryotes harbor similar genetic clusters for the synthesis of thiazole and (methyl)-oxazole heterocycles (3, 18, 19). Additional important mammalian pathogens such as Listeria monocytogenes, Staphylococcus aureus, and Clostridium botulinum, contain sag-like gene clusters that produce SLS-like cytolysins. These toxins are expected to promote pathogen survival and host cell injury during infection, but this has only been conclusively shown for S. pyogenes and L. monocytogenes (2, 18). Like E. coli, many other prokaryotes harbor a sag-like genetic cluster but are not known to produce cytolysins. Some examples are the goadsporin-producing organism, Streptomyces sp. TP-A0584 and cyanobactin producers such as Prochloron didemni (2022). The molecular targets of these secondary metabolites remain to be elucidated, but it is known that goadsporin exhibits antibiotic activity, and the cyanobactin, patellamide D, reverses multiple drug resistance in a human leukemia cell line (23). Because genetic loci containing sagBCD-like genes have been widely disseminated in prokaryotes (3), nature appears to have found a preferred route to synthesizing such secondary metabolites.In this work, we build upon our initial report on the in vitro reconstitution of SLS biosynthesis to uncover the requisite features of substrate selectivity and cytolytic activity. The impetus for defining substrate tolerance arose from earlier results showing that SagBCD accepts alternate substrates in vitro (3), as evidenced by two key experiments. First, SagBCD converted a noncognate substrate, ClosA (C. botulinum), into a cytolytic entity. Second, mass spectrometry revealed heterocycle formation on the McbA (E. coli) peptide after SagBCD treatment (3). Here, we dissect the N-terminal leader peptide and C-terminal protoxin of SagA to define the residues necessary for conversion into SLS.  相似文献   

19.
Glucosinolates are plant secondary metabolites present in Brassicaceae plants such as the model plant Arabidopsis thaliana. Intact glucosinolates are believed to be biologically inactive, whereas degradation products after hydrolysis have multiple roles in growth regulation and defense. The degradation of glucosinolates is catalyzed by thioglucosidases called myrosinases and leads by default to the formation of isothiocyanates. The interaction of a protein called epithiospecifier protein (ESP) with myrosinase diverts the reaction toward the production of epithionitriles or nitriles depending on the glucosinolate structure. Here we report the identification of a new group of nitrile-specifier proteins (AtNSPs) in A. thaliana able to generate nitriles in conjunction with myrosinase and a more detailed characterization of one member (AtNSP2). Recombinant AtNSP2 expressed in Escherichia coli was used to test its impact on the outcome of glucosinolate hydrolysis using a gas chromatography-mass spectrometry approach. AtNSP proteins share 30–45% sequence homology with A. thaliana ESP. Although AtESP and AtNSP proteins can switch myrosinase-catalyzed degradation of 2-propenylglucosinolate from isothiocyanate to nitrile, only AtESP generates the corresponding epithionitrile. Using the aromatic benzylglucosinolate, recombinant AtNSP2 is also able to direct product formation to the nitrile. Analysis of glucosinolate hydrolysis profiles of transgenic A. thaliana plants overexpressing AtNSP2 confirms its nitrile-specifier activity in planta. In silico expression analysis reveals distinctive expression patterns of AtNSPs, which supports a biological role for these proteins. In conclusion, we show that AtNSPs belonging to a new family of A. thaliana proteins structurally related to AtESP divert product formation from myrosinase-catalyzed glucosinolate hydrolysis and, thereby, likely affect the biological consequences of glucosinolate degradation. We discuss similarities and properties of AtNSPs and related proteins and the biological implications.Brassicaceae plants such as oilseed rape (Brassica napus), turnip (Brassica rapa), and white mustard (Sinapis alba) as well as the model plant Arabidopsis thaliana contain a group of secondary metabolites known as glucosinolates (GSLs)2 (1, 2). These are β-thioglucoside N-hydroxysulfates with a sulfur-linked β-d-glucopyranose moiety and a variable side chain that is derived from one of eight amino acids or their methylene group-elongated derivatives. Aliphatic GSLs are derived from alanine, leucine, isoleucine, valine, or predominantly methionine. Tyrosine or phenylalanine give aromatic GSLs, and tryptophan-derived GSLs are called indolic GSLs (for review, see Ref. 3). Although more than 120 different GSLs have been identified in total so far, individual plant species usually contain only a few GSLs (2). Quantitative and qualitative differences of GSL profiles are also observed within a species, such as, for example, for different A. thaliana ecotypes (46). In addition, GSL composition varies among organs and during the life cycle of plants (7, 8) and is affected by external factors (9).Intact GSLs are mostly considered to be biologically inactive. Most GSL degradation products have toxic effects on insect, fungal, and bacterial pests, serve as attractants for specialist insects, or may have beneficial health effects for humans (1015). The enzymatic degradation of GSLs (Fig. 1A), which occurs massively upon tissue damage, is catalyzed by plant thioglucosidases called myrosinases (EC 3.2.1.147; glycoside hydrolase family 1). Depending on several factors (e.g. GSL structure, proteins, cofactors, pH) myrosinase-catalyzed hydrolysis of GSLs can lead to a variety of products (Fig. 1B; for review, see Refs. 16 and 17). Of these, isothiocyanates are the most common as their formation only requires myrosinase activity. Thiocyanates on the other hand are only produced from a very limited number of GSLs, and their formation necessitates the presence of a thiocyanate-forming factor in addition to myrosinase (18). A thiocyanate-forming protein (TFP) has recently been identified in Lepidium sativum (19). Alkenyl GSLs, a subgroup of aliphatic GSLs containing a terminal unsaturation in their side chain, can lead to the production of epithionitriles through the cooperative action of myrosinase and a protein called epithiospecifier protein (ESP (20)) in a ferrous ion-dependent way (2123). Both TFP and ESP contain a series of Kelch repeats (19). Kelch repeats are involved in protein-protein interactions, and Kelch repeat-containing proteins are involved in a number of diverse biological processes (24). In addition to isothiocyanates, nitriles are the major group of GSL hydrolysis products. Although ESP and TFP activities can generate nitriles (19, 21, 25, 26), indications for an ESP-independent nitrile-specifier activity exist. The GSL hydrolysis profile of A. thaliana roots, an organ that does not show ESP expression or activity (27), reveals predominantly the presence of nitriles (28). In addition, leaf tissue of A. thaliana ecotypes supposedly devoid of ESP activity produces a certain amount of nitriles upon autolysis (21). Under acidic buffer conditions, a non-enzymatic production of nitriles from GSLs is observed (Ref. 29 and references therein). Increasing Fe2+ concentrations have also been shown to favor nitrile formation over isothiocyanate formation from a number of GSLs in the presence of myrosinase and absence of ESP (21, 22). Therefore, a non-enzymatic origin of this nitrile production cannot be excluded, although the presence of a nitrile-specifier protein is a tempting alternative. Although ESP is able to generate nitriles, it has also been shown that the conversion rates of GSLs to nitriles are lower than those of GSLs to epithionitriles for ESP (21, 22).Open in a separate windowFIGURE 1.Simplified scheme of enzymatic GSL hydrolysis (A) and structures and names of GSLs and their hydrolysis products that are mentioned in the article. (B). A, myrosinase acts on GSLs to form an unstable aglycone intermediate that can rearrange spontaneously to form an isothiocyanate. Hydrolysis can be diverted from this default route under certain conditions (e.g. the presence of NSPs, ferrous ions, or at pH < 5) to give the corresponding nitrile. ESP is responsible for the formation of epithionitriles from alkenyl GSLs in a ferrous ion-dependent mechanism. B, the general structure of GSLs, indicating the variable side chain as R, is given as well as the three major classes of hydrolysis products (i.e. isothiocyanates, nitriles, and epithionitriles). The listed GSLs are the ones mentioned in this article and are arranged according to the class of GSLs they belong to and with an increase in chain length or complexity. The names of the respective hydrolysis products are given for a better understanding of the present article, and not all were encountered during our studies.A nitrile-specifier protein (NSP) that is able to redirect the hydrolysis of GSLs toward nitriles has been cloned from the larvae of the butterfly Pieris rapae (30). This protein does not, however, exhibit sequence similarity to plant ESP, and a corresponding plant nitrile-specifier protein has not yet been identified. We report here the identification of a group of six A. thaliana genes with some sequence similarity to A. thaliana ESP, providing evidence for a new family of nitrile-specifier proteins and a more detailed characterization of one member that possesses nitrile-specifier activity in vitro, when applied exogenously to plant tissue and after ectopic expression in the two A. thaliana ecotypes Col-0 and C24. Despite its sequence homology to A. thaliana epithiospecifier protein (AtESP), it does not possess epithiospecifier activity under similar conditions. Therefore, we propose to designate this protein as A. thaliana nitrile-specifier protein 2 (AtNSP2). Although the biological roles of AtNSP2 and related proteins are not yet known, their specificities and distinctive expression patterns indicate the presence of a fine-tuned mechanism for GSL degradation controlling the outcome of an array of biologically active molecules.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号