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 140 毫秒
1.
Rab proteins influence vesicle trafficking pathways through the assembly of regulatory protein complexes. Previous investigations have documented that Rab11a and Rab8a can interact with the tail region of myosin Vb and regulate distinct trafficking pathways. We have now determined that a related Rab protein, Rab10, can interact with myosin Va, myosin Vb, and myosin Vc. Rab10 localized to a system of tubules and vesicles that have partially overlapping localization with Rab8a. Both Rab8a and Rab10 were mislocalized by the expression of dominant-negative myosin V tails. Interaction with Rab10 was dependent on the presence of the alternatively spliced exon D in myosin Va and myosin Vb and the homologous region in myosin Vc. Yeast two-hybrid assays and fluorescence resonance energy transfer studies confirmed that Rab10 binding to myosin V tails in vivo required the alternatively spliced exon D. In contrast to our previous work, we found that Rab11a can interact with both myosin Va and myosin Vb tails independent of their splice isoform. These results indicate that Rab GTPases regulate diverse endocytic trafficking pathways through recruitment of multiple myosin V isoforms.Eukaryotic cells are comprised of networks of highly organized membranous structures that require the efficient and timely movement of diverse intracellular proteins for proper function. Molecular motors provide the physical force needed to move these materials along microtubules and actin microfilaments. Unconventional myosin motors, such as those belonging to classes V, VI, and VII, have roles in the trafficking and recycling of membrane-bound structures in eukaryotic cells (1) and are recruited to discrete vesicle populations. Myosin VI is involved in clathrin-mediated endocytosis (2), whereas myosin VIIa participates in the proper development of stereocilia of inner ear hair cells and the transport of pigment granules in retinal pigmented epithelial cells (3, 4). Similarly, the three members of vertebrate class V myosins, myosin Va, myosin Vb, and myosin Vc, are required for the proper transport of a wide array of membrane cargoes, such as the melanosomes of pigment cells, synaptic vesicles in neurons, apical recycling endosomes in polarized epithelial cells, and bulk recycling vesicles in non-polarized cells (5).Members of the Rab family of small GTPases regulate many cellular systems, including membrane trafficking (6, 7). Certain Rab proteins associate with and regulate the function of class V myosins. Rab27a, in a complex with the adaptor protein melanophilin/Slac2-a, is required to localize myosin Va to the surface of melanin-filled pigment granules in vertebrates (8-10), whereas Rab27a and Slac2-c/MyRIP associate with both Myosin Va and myosin VIIa (3, 11). Rab11a, in a complex with its adaptor protein Rab11-FIP2, associates with myosin Vb on recycling endosomes (12-14) where the tripartite complex regulates the recycling of a variety of cargoes (15-19). In addition, Rab8a associates with both myosin Vb (20) and myosin Vc (21) as part of the non-clathrin-mediated tubular recycling system (20). Recently, Rab11a has also been shown to associate with myosin Va in the transport of AMPA receptors in dendritic spines (22), contributing to the model of myosin V regulation by multiple Rab proteins.Previous investigations have documented alternative splicing of myosin Va in a tissue-specific manner (23-28). Alternate splicing occurs in a region lying between the coiled-coil region of the neck of the motor and the globular tail region. Three exons in particular are subject to alternative splicing: exons B, D, and F (23-25). Exon F is critical for association with melanophilin/Slac2 and Rab27a (8, 9, 29, 30). Additionally, exon B is required for the interaction of myosin Va with dynein light chain 2 (DLC2) (27, 28). Currently no function for the alternatively spliced exon D has been reported. Similar to myosin Va, myosin Vb contains exons A, B, C, D, and E, whereas no exon F has yet been identified in myosin Vb (Fig. 1A). In addition, exon B in myosin Vb does not resemble the dynein light chain 2 (DLC2) binding region in myosin Va (27, 28), and therefore, it likely does not interact with DLC2. On the other hand, exon D is highly conserved among Myosin Va, myosin Vb, and myosin Vc, suggesting a common function in these molecular motors.Open in a separate windowFIGURE 1.Tissue distribution of human myosin Va and myosin Vb splice isoforms. A, schematic of the alternative exon organization in the tails of myosin Va and myosin Vb. It is known that exons B, D, and F are subject to alternative splicing in myosin Va, whereas there is only evidence that exon D is alternatively spliced in myosin Vb, which does not contain exon F. B, alignment of exon D sequences from mouse and human myosin V''s. myosin Va and myosin Vb both contain exon D (amino acids 1320-1346 of myosin Va and 1315-1340 of myosin Vb), whereas myosin Vc contains an exon D-like region (amino acids 1124-1147 of human myosin Vc) that is not known to be alternatively spliced. Alignment of the exon D regions from all three motors reveals a high degree of homology, especially in the center of the exon. Asterisks indicate amino acid identities. C, PCR-based analysis of human tissue panels reveals the alternative splicing pattern of exon D in myosin Va and myosin Vb. Primers flanking the region encoding exon D for both motors were used to amplify cDNA from human MTC™ panels (Clontech). cDNA amplified from HeLa cell RNA as well as myosin Va and myosin Vb tail constructs were used as positive controls. Variants expressing exon D (upper bands) and lacking exon D (lower bands) were visible. Per., peripheral; Pos., positive.Here we report that Rab10, a protein related to Rab8a and thought to have similar function (31-35), localizes to a system of tubules and vesicles overlapping in distribution with Rab8a in HeLa cells. Utilizing dominant-negative myosin V tail constructs, we show that Rab8a and Rab10 can interact with Myosin Va, myosin Vb, and myosin Vc in vivo. In addition, we have determined that the alternatively spliced exon D in both myosin Va and myosin Vb is required for interaction with Rab10. In contrast to our previous findings, we demonstrate that Rab11a is able to interact with both myosin Va and myosin Vb tails in an exon independent-manner. These results reveal that multiple Rab proteins potentially regulate all three class V myosin motors.  相似文献   

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

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

4.
5.
The nonmuscle myosin IIA heavy chain (Myh9) is strongly associated with adhesion structures of osteoclasts. In this study, we demonstrate that during osteoclastogenesis, myosin IIA heavy chain levels are temporarily suppressed, an event that stimulates the onset of cell fusion. This suppression is not mediated by changes in mRNA or translational levels but instead is due to a temporary increase in the rate of myosin IIA degradation. Intracellular activity of cathepsin B is significantly enhanced at the onset of osteoclast precursor fusion, and specific inhibition of its activity prevents myosin IIA degradation. Further, treatment of normal cells with cathepsin B inhibitors during the differentiation process reduces cell fusion and bone resorption capacity, whereas overexpression of cathepsin B enhances fusion. Ongoing suppression of the myosin IIA heavy chain via RNA interference results in formation of large osteoclasts with significantly increased numbers of nuclei, whereas overexpression of myosin IIA results in less osteoclast fusion. Increased multinucleation caused by myosin IIA suppression does not require RANKL. Further, knockdown of myosin IIA enhances cell spreading and lessens motility. These data taken together strongly suggest that base-line expression of nonmuscle myosin IIA inhibits osteoclast precursor fusion and that a temporary, cathepsin B-mediated decrease in myosin IIA levels triggers precursor fusion during osteoclastogenesis.The final stages of osteoclastogenesis involve fusion of differentiated precursors from the monocyte/macrophage lineage (1). Although the membrane structural components regulating preosteoclast fusion are not well understood, in recent years a number of candidate cell surface molecules have been implicated, including receptors CD44 (2, 3), CD47 and its ligand macrophage fusion receptor (also known as signal regulatory protein α) (46), the purinergic receptor P2X7 (7), and the disintegrin and metalloproteinase ADAM8 (8). A recently identified receptor, the dendritic cell-specific transmembrane protein, is essential for osteoclast fusion both in vitro and in vivo (9, 10). More recently, the d2 subunit of proton-translocating vacuolar proton-translocating ATPases, a membrane subunit isoform expressed predominantly in osteoclasts, similarly was demonstrated to be required for fusion in vitro and in vivo (11). However, elucidation of the mechanisms by which these molecules may mediate cell fusion has proved to be difficult.The mammalian class II myosin family consists of distinct isoforms expressed in skeletal, smooth, and cardiac muscle, as well as three nonmuscle forms designated IIA, IIB, and IIC (1214). Although all class II molecules are composed of two heavy chains, two essential light chains, and two regulatory chains, their unique activities are a function of their particular heavy chain isoforms. Although the nonmuscle heavy chain isoforms share extensive structural homology, they have been shown to demonstrate distinct patterns of expression (1518), enzyme kinetics and activation (12, 1921), and cellular function (2224). Knock-out of either myosin IIA or IIB results in embryonic lethality, although death derives from defects unique to each isoform (25, 26). In vitro, myosin IIA, a target of Rho kinase, has been shown to be involved in a wide variety of cellular functions, including cytokinesis, cell contractility, and adhesion and motility.The actin cytoskeleton of osteoclasts possesses features unlike those of most mammalian cell types. First, osteoclasts do not possess stress fibers but instead form a meshwork of fine actin filaments throughout the cell (2729). Osteoclasts express unusual attachment structures typified by the podosome, a form of adhesion structure most typically present in cells of the monocyte/macrophage lineage, dendritic cells, and smooth muscle cells. Podosomes are integrin-based cell-matrix contact structures that are notable for the presence of a short (0.5–1.0 μm) F-actin core surrounded by a ring of adaptor proteins, kinases, small GTPases, and regulators of endocytosis (30, 31). When cultured on glass, mature osteoclasts generate a belt of podosomes at the cell periphery. However, when cultured on bone, osteoclasts form a dense ring of podosome-like structures that is usually internal to the cell margins (32). This region, termed the sealing zone, surrounds a specialized membrane domain termed the ruffled border, from which protons and proteases are secreted to induce resorption of bone (1). We previously demonstrated that myosins IIA and IIB localize to distinct subcellular regions within osteoclasts, with MyoIIA2 strongly segregating to both podosomes and the actin ring of the sealing zone (28). Because of this distribution into osteoclast adhesion structures and findings in other cells showing MyoIIA to be associated with dynamic Rho-kinase-dependent functions, such as adhesion and locomotion, we hypothesized that MyoIIA may play a vital role in cell motility and the bone resorption function. In this study, we examined cellular expression of MyoIIA during osteoclastogenesis and, along with RNA interference-mediated suppression of the protein, have confirmed its role in cell spreading, motility, and sealing zone formation. However, this study also unexpectedly revealed a role for MyoIIA in regulating preosteoclast fusion during osteoclastogenesis.  相似文献   

6.
7.
8.
9.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

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

11.
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   

12.
13.
Bile acids are steroid detergents that are toxic to mammalian cells at high concentrations; increased exposure to these steroids is pertinent in the pathogenesis of cholestatic disease and colon cancer. Understanding the mechanisms of bile acid toxicity and apoptosis, which could include nonspecific detergent effects and/or specific receptor activation, has potential therapeutic significance. In this report we investigate the ability of synthetic enantiomers of lithocholic acid (ent-LCA), chenodeoxycholic acid (ent-CDCA), and deoxycholic acid (ent-DCA) to induce toxicity and apoptosis in HT-29 and HCT-116 cells. Natural bile acids were found to induce more apoptotic nuclear morphology, cause increased cellular detachment, and lead to greater capase-3 and -9 cleavage compared with enantiomeric bile acids in both cell lines. In contrast, natural and enantiomeric bile acids showed similar effects on cellular proliferation. These data show that bile acid-induced apoptosis in HT-29 and HCT-116 cells is enantiospecific, hence correlated with the absolute configuration of the bile steroid rather than its detergent properties. The mechanism of LCA- and ent-LCA-induced apoptosis was also investigated in HT-29 and HCT-116 cells. These bile acids differentially activate initiator caspases-2 and -8 and induce cleavage of full-length Bid. LCA and ent-LCA mediated apoptosis was inhibited by both pan-caspase and selective caspase-8 inhibitors, whereas a selective caspase-2 inhibitor provided no protection. LCA also induced increased CD95 localization to the plasma membrane and generated increased reactive oxygen species compared with ent-LCA. This suggests that LCA/ent-LCA induce apoptosis enantioselectively through CD95 activation, likely because of increased reactive oxygen species generation, with resulting procaspase-8 cleavage.Bile acids are physiologic steroids that are necessary for the proper absorption of fats and fat-soluble vitamins. Their ability to aid in these processes is largely due to their amphipathic nature and thus their ability to act as detergents. Despite the beneficial effects, high concentrations of bile acids are toxic to cells (1-11). High fat western diets induce extensive recirculation of the bile acid pool, resulting in increased exposure of the colonic epithelial cells to these toxic steroids (12, 13). A high fat diet is also a risk factor for colon carcinogenesis; increased bile acid exposure is responsible for some of this risk. Bile acids can contribute to both colon cancer formation and progression, and their effects on colonic proliferation and apoptosis aid this process by disrupting the balance between cell growth and cell death, as well as helping to select for bile acid-resistant cells (14, 15).In colonocyte-derived cell lines bile acid-induced apoptosis is thought to proceed through mitochondrial destabilization with resulting mitochondrial permeability transition formation and cytochrome c release as well as generation of oxidative stress (1, 9-11). Bile acid-induced apoptosis has also been extensively explored in hepatocyte derived cell lines with mechanisms including mitochondria dysfunction (16-23), endoplasmic reticulum stress (24), ligand-independent activation of death receptor pathways (18, 25-28), and modulation of cellular apoptotic and anti-apoptotic Bcl-2 family proteins (29).Although ample evidence exists for multiple mechanisms of bile acid-induced apoptosis, the precise interactions responsible for initiating these apoptotic pathways are still unclear. Bile acids have been shown to interact directly with specific receptors (30, 31). These steroids can also initiate cellular signaling through nonspecific membrane perturbations (32), and evidence exists showing that other simple detergents (i.e. Triton X-100) are capable of inducing caspase cleavage nonspecifically with resultant apoptosis (33). Therefore, hydrophobic bile acids may interact nonspecifically with cell membranes to alter their physical properties, bind to receptors specific for these steroids, or utilize a combination of both specific and nonspecific interactions to induce apoptosis.Bile acid enantiomers could be useful tools for elucidating mechanisms of bile acid toxicity and apoptosis. These enantiomers, known as ent-bile acids, are synthetic nonsuperimposable mirror images of natural bile acids with identical physical properties except for optical rotation. Because bile acids are only made in one absolute configuration naturally, ent-bile acids must be constructed using a total synthetic approach. Recently we reported the first synthesis of three enantiomeric bile acids: ent-lithocholic acid (ent-LCA),2 ent-chenodeoxycholic acid (ent-CDCA), and ent-deoxycholic acid (ent-DCA) (Fig. 1) (34, 35). Enantiomeric bile acids have unique farnesoid X receptor, vitamin D receptor, pregnane X receptor, and TGR5 receptor activation profiles compared with the corresponding natural bile acids (34). This illustrates that natural and enantiomeric bile acids interact differently within chiral environments because of their distinct three-dimensional configurations (Fig. 1). Despite these differences in chiral interactions, ent-bile acids have physical properties identical to those of their natural counterparts including solubility and critical micelle concentrations (34, 35). With different receptor interaction profiles and identical physical properties compared with natural bile acids, ent-bile acids are ideal compounds to differentiate between the receptor-mediated and the non-receptor-mediated functions of natural bile acids.Open in a separate windowFIGURE 1.Natural and enantiomeric bile acids. Structures and three-dimensional projection views of natural LCA, CDCA, DCA, and their enantiomers (ent-LCA, ent-CDCA, and ent-DCA). The three-dimensional ent-steroid structure is rotated 180° around the long axis for easier comparison with the natural steroid.In this study we explore the enantioselectivity of LCA-, CDCA-, and DCA-mediated toxicity and apoptosis in two human colon adenocarcinoma cell lines, HT-29 and HCT-116. Because the mechanism of natural LCA induced apoptosis has never been characterized, we then examined in more detail LCA- and ent-LCA-mediated apoptosis in colon cancer cells. These studies will not only explore the LCA apoptotic mechanism but will also determine whether ent-LCA signals through similar cellular pathways.  相似文献   

14.
The adhesion of bacteria to host tissues is often mediated by interactions with extracellular matrices. Herein, we report on the interactions of the group A streptococcus, Streptococcus pyogenes, with the extracellular matrix protein fibulin-1. S. pyogenes bound purified fibulin-1 in a dose-dependent manner. Genetic ablation of serum opacity factor (SOF), a virulence determinant of S. pyogenes, reduced binding by ∼50%, and a recombinant peptide of SOF inhibited binding of fibulin-1 to streptococci by ∼45%. Fibulin-1 bound to purified SOF2 in a dose-dependent manner with high affinity (Kd = 1.6 nm). The fibulin-1-binding domain was localized to amino acid residues 457–806 of SOF2, whereas the fibronectin-binding domain is contained within residues 807–931 of SOF2, indicating that these two domains are separate and distinct. Fibulin-1 bound to recombinant SOF from M types 2, 4, 28, and 75 of S. pyogenes, indicating that the fibulin-1-binding domain is likely conserved among SOF from different serotypes. Mixed binding experiments suggested that gelatin, fibronectin, fibulin-1, and SOF form a quaternary molecular complex that enhanced the binding of fibulin-1. These data indicate that S. pyogenes can interact with fibulin-1 and that SOF is a major streptococcal receptor for fibulin-1 but not the only receptor. Such interactions with fibulin-1 may be involved in the adhesion of S. pyogenes to extracellular matrices of the host.Adhesion of bacteria to host surfaces is the first stage in establishing bacterial infections in the human host, and a variety of molecular mechanisms are utilized to initiate adhesion. A common mechanism for adhesion involves interactions between bacterial adhesins and components of the extracellular matrices of the host. The identification and characterization of microbial surface components recognizing adhesive matrix molecules (MSCRAMM) has led to important advances in vaccines and immunotherapies for preventing and treating bacterial infections (1).The group A streptococcus, Streptococcus pyogenes, is a major human pathogen causing diseases ranging from relative minor infections such as pharyngitis and cellulitis to severe infections with high levels of morbidity and mortality such as necrotizing fasciitis and toxic shock syndrome (2). This pathogen expresses adhesins that interact with various components of the extracellular matrix including laminin, elastin, fibronectin, fibrinogen, and collagen (37). The interactions between fibronectin and S. pyogenes have been intensely studied, and these investigations have revealed at least 10 different streptococcal proteins that bind fibronectin (4).Serum opacity factor (SOF)2 is a major fibronectin-binding protein that is involved in adhesion to host cells (811). SOF is a virulence determinant that is expressed by approximately half of the clinical isolates of S. pyogenes (8). SOF opacifies serum by binding and displacing apoA-I in high density lipoproteins (8, 1215). SOF is covalently linked to the streptococcal cell wall via an LPSTG sortase recognition site and is also released in a soluble form. SOF has two functionally distinct domains, an N-terminal domain that opacifies serum and a C-terminal domain that binds fibronectin. The role of SOF in adhesion involves both its C-terminal fibronectin-binding domain and an N-terminal region (see Fig. 1 for a schematic of structure) (9, 11). However, the nature of the interactions between the N-terminal region of SOF and host components is not well characterized.Open in a separate windowFIGURE 1.A, a schematic of the structure of SOF and its functional domains is shown. The assignment of functional domains are based on the findings of Rakonjac et al. (33), Kreikemeyer et al. (34), Courtney et al. (8, 13), and results presented in this work. Fn, fibronectin. B, the data for the binding of SOF peptides to fibronectin are from previous publications (8, 13), and the data for fibulin-1 are from the present work.Herein, we report on the interactions between a truncated form of SOF in which its fibronectin-binding domain has been deleted and the extracellular matrix protein fibulin-1. Fibulin-1 is a member of the fibulin family that currently consists of seven glycoproteins. All fibulins contain epidermal growth factor-like repeats and a unique fibulin-type module at its C terminus that define this family (16, 17). Fibulin-1 is found within the extracellular matrices and in human plasma at 30–50 μg/ml (18). It interacts with many of the components of the extracellular matrix including fibronectin, laminin, fibrinogen, nidogen-1, endostatin, aggrecan, and versican (16, 19). Due to its intimate relationship with the extracellular matrix, it is not surprising that the defects in fibulin-1 have a wide-ranging impact. Genetic evidence suggests that fibulin-1 is involved in tissue organization, the maturation and maintenance of blood vessels, and multiple embryonic pathways (16, 2022).Although it has been established that many of the other components of the extracellular matrix can interact with bacteria, there has been no previous report on the binding of fibulins to bacteria. Our findings indicate that fibulin-1 does bind to streptococci and that SOF is a major streptococcal receptor for fibulin-1.  相似文献   

15.
Cardiac myosin-binding protein C (cMyBP-C) is a regulatory protein expressed in cardiac sarcomeres that is known to interact with myosin, titin, and actin. cMyBP-C modulates actomyosin interactions in a phosphorylation-dependent way, but it is unclear whether interactions with myosin, titin, or actin are required for these effects. Here we show using cosedimentation binding assays, that the 4 N-terminal domains of murine cMyBP-C (i.e. C0-C1-m-C2) bind to F-actin with a dissociation constant (Kd) of ∼10 μm and a molar binding ratio (Bmax) near 1.0, indicating 1:1 (mol/mol) binding to actin. Electron microscopy and light scattering analyses show that these domains cross-link F-actin filaments, implying multiple sites of interaction with actin. Phosphorylation of the MyBP-C regulatory motif, or m-domain, reduced binding to actin (reduced Bmax) and eliminated actin cross-linking. These results suggest that the N terminus of cMyBP-C interacts with F-actin through multiple distinct binding sites and that binding at one or more sites is reduced by phosphorylation. Reversible interactions with actin could contribute to effects of cMyBP-C to increase cross-bridge cycling.Cardiac myosin-binding protein C (cMyBP-C)2 is a thick filament accessory protein that performs both structural and regulatory functions within vertebrate sarcomeres. Both roles are likely to be essential in deciphering how a growing number of mutations found in the cMyBP-C gene, i.e. MYBPC3, lead to cardiomyopathies and heart failure in a substantial number of the world''s population (1, 2).Considerable progress has recently been made in determining the regulatory functions of cMyBP-C and it is now apparent that cMyBP-C normally limits cross-bridge cycling kinetics and is critical for cardiac function (3-5). Phosphorylation of cMyBP-C is essential for its regulatory effects because elimination of phosphorylation sites (serine to alanine substitutions) abolishes the ability of protein kinase A (PKA) to accelerate cross-bridge cycling kinetics and blunts cardiac responses to inotropic stimuli (6). The substitutions further impair cardiac function, reduce contractile reserve, and cause cardiac hypertrophy in transgenic mice (6, 7). By contrast, substitution of aspartic acids at these sites to mimic constitutive phosphorylation is benign or cardioprotective (8).Although a role for cMyBP-C in modulating cross-bridge kinetics is supported by several transgenic and knock-out mouse models (6, 7, 9, 10), the precise mechanisms by which cMyBP-C exerts these effects are not completely understood. For instance, the unique regulatory motif or “m-domain” of cMyBP-C binds to the S2 subfragment of myosin in vitro (11) and binding is abolished by PKA-mediated phosphorylation of the m-domain (12). These observations have led to the idea that (un)binding of the m-domain from myosin S2 mediates PKA-induced increases in cross-bridge cycling kinetics. Consistent with this idea, Calaghan and colleagues (13) showed that S2 added to transiently permeabilized myocytes increased their contractility, presumably because added S2 displaced cMyBP-C from binding endogenous S2. However, other reports indicate that cMyBP-C can influence actomyosin interactions through mechanisms unrelated to S2 binding, because either purified cMyBP-C (14) or recombinant N-terminal domains of cMyBP-C (15) affected acto-S1 filament sliding velocities and ATPase rates in the absence of myosin S2. These results thus raise the possibility that interactions with ligands other than myosin S2, such as actin or myosin S1, contribute to effects of cMyBP-C on cross-bridge interaction kinetics.The idea that cMyBP-C interacts with actin to influence cross-bridge cycling kinetics is supported by several studies that implicate the regulatory m-domain or sequences near it in actin binding (16-19). cMyBP-C is a member of the immunoglobulin (Ig) superfamily of proteins and consists of 11 repeating domains that bear homology to either Ig or fibronectin-like folds. Domains are numbered sequentially from the N terminus of cMyBP-C as C0 through C10. The m-domain, a unique sequence of ∼100 amino acids, is located between domains C1 and C2 and is phosphorylated on at least 3 serine residues by PKA (12). Although the precise structure of the m-domain is not known, small angle x-ray scattering data suggest that it is compact and folded in solution and is thus similar in size and dimensions to the surrounding Ig domains (20). Recombinant proteins encompassing the m-domain and/or a combination of adjacent domains including C0, C1, C2, and a proline-alanine-rich sequence that links C0 to C1 have been shown to bind actin (16, 18, 19).The purpose of the present study was to characterize binding interactions of the N terminus of cMyBP-C with actin and to determine whether interactions with actin are influenced by phosphorylation of the m-domain. Results demonstrate that the N terminus of cMyBP-C binds to F-actin and to native thin filaments with affinities similar to that reported for cMyBP-C binding to myosin S2 (11). Furthermore, actin binding was reduced by m-domain phosphorylation, suggesting that reversible interactions of cMyBP-C with actin could contribute to modulation of cross-bridge kinetics.  相似文献   

16.
Two thioesterases are commonly found in natural product biosynthetic clusters, a type I thioesterase that is responsible for removing the final product from the biosynthetic complex and a type II thioesterase that is believed to perform housekeeping functions such as removing aberrant units from carrier domains. We present the crystal structure and the kinetic analysis of RifR, a type II thioesterase from the hybrid nonribosomal peptide synthetases/polyketide synthase rifamycin biosynthetic cluster of Amycolatopsis mediterranei. Steady-state kinetics show that RifR has a preference for the hydrolysis of acyl units from the phosphopantetheinyl arm of the acyl carrier domain over the hydrolysis of acyl units from the phosphopantetheinyl arm of acyl-CoAs as well as a modest preference for the decarboxylated substrate mimics acetyl-CoA and propionyl-CoA over malonyl-CoA and methylmalonyl-CoA. Multiple RifR conformations and structural similarities to other thioesterases suggest that movement of a helical lid controls access of substrates to the active site of RifR.Assembly line complexes, which include modular polyketide synthases (PKS)3 and nonribosomal peptide synthetases (NRPS), are multifunctional proteins composed of modules that work in succession to synthesize secondary metabolites, many of which are precursors of potent antibiotics, immunosuppressants, anti-tumor agents, and other bioactive compounds. Rifamycin, the precursor to the anti-tuberculosis drug rifampicin, is produced by the rifamycin assembly line complex, which is an NRPS/PKS hybrid system composed of one NRPS-like and 10 PKS modules (1). Each module in an assembly line complex extends and modifies the intermediate compound before passing it on to the next module in the series (Fig. 1A). The intermediate compounds are covalently attached through a thioester linkage to the phosphopantetheine arm (Ppant) of carrier domains, one associated with each module, until they are released from the synthase, usually by a type I thioesterase (TEI) (2, 3).Open in a separate windowFIGURE 1.Proposed functions of thioesterase II proteins. A, chain elongation by a PKS module. The chain elongation intermediate is transferred from the ACP of the upstream module to the ketosynthase (KS) domain. The acyltransferase (AT) domain transfers an acyl group building block from CoA to the ACP within the module. The KS domain catalyzes condensation of the new building block with the intermediate, releasing CO2. B, production of a decarboxylated acyl unit by the ketosynthase domain and the subsequent hydrolysis by a TEII. C, mispriming of a PKS by transfer of an acyl-phosphopantetheine arm by a promiscuous phosphopantetheinyl transferase (Pptase) and the subsequent hydrolysis by a TEII. D, hydrolysis of an amino acid derivative by a TEII from an NRPS module comprising an adenylation domain (A) and a peptide carrier protein (PCP) domain.TEIs are usually integrated into the final module of the assembly line complex and remove the final product through macrocyclization or hydrolysis. Occasionally, tandem type I thioesterases are integrated at the C terminus of the final module of NRPS pathways (4).Although TEIs are covalently attached to the terminal module and generally process only the final product of an assembly line complex, type II thioesterases (TEIIs) are discrete proteins that can remove intermediates from any module in the complex. A variety of functions have been attributed to TEIIs, the most prevalent of which is a “housekeeping function,” the removal of aberrant acyl units from carrier domains. These aberrant acyl units may be due to premature decarboxylation by a PKS ketosynthase domain (5) (Fig. 1B) or to mispriming of the carrier domain by a promiscuous phosphopantetheinyl transferase (68) (Fig. 1C). Other proposed functions for TEIIs include the removal of intermediates from the synthase as in the case of the mammary gland rat fatty acid synthase (FAS) TEII in lactating rats, which removes medium chain C8-C12 fatty acids from the ACP domain (9) and the removal of amino acid derivatives from a carrier domain (1013), allowing these derivatives to be incorporated into the natural product by a later module in the assembly line complex (Fig. 1D).Disruption of the TEI function results in a complete loss of product, whereas disruption of TEII function results in a significant decrease in product yield (30–95%) (4, 1424). Removal of the TEII from the rifamycin assembly line resulted in a 60% decrease in product yield (25). Neither TEIs nor TEIIs may rescue the disrupted function of the other (6), but a TEII from another pathway may rescue the function of a disrupted TEII (26).Two models have been proposed for the TEII housekeeping function (5). In the high specificity model, the TEII scans the complex and efficiently removes only aberrant acyl units. In the low specificity model, the TEII removes both correct and incorrect acyl units from the Ppant arm at an inefficient rate. Correct acyl units are quickly incorporated into the growing intermediate compound. In contrast, incorrect acyl units stall the assembly line, providing a longer window of opportunity for removal by a TEII. Thus a slow, low specificity enzyme can be effective.TEIIs from different pathways have differing specificities, but general trends include a preference for decarboxylated acyl units over carboxylated acyl units (5, 6, 27), substrates linked to a carrier domain over substrates linked to CoA or the phosphopantetheine mimic N-acetylcysteamine (7, 28), and single amino acids over di- or tri-peptides (6, 7). TEIIs are able to hydrolyze substrates attached to carrier domains from their native pathway as well as other pathways (6, 20, 28).PKS/NRPS/FAS thioesterases belong to the α/β hydrolase family. Structures are reported for seven PKS/NRPS/FAS thioesterases: crystal structures for the TEIs from the pikromycin (PikTE) PKS (29), 6-deoxyerythronolide B (DEBSTE) PKS (30), surfactin NRPS (SrfTE) (31), fengycin NRPS (FenTE) (32), and human fatty acid synthase (hFasTE) (33) systems, and NMR structures for enterobactin TEI (34), and surfactin TEII (35). Like PKS modules, PKS TEIs are dimers. The dimer interface comprises two N-terminal helices that are unique to the PKS TEIs. NRPS TEIs are monomeric, like NRPS modules. The NRPS TEII of surfactin is also monomeric (31). Although the FAS complex is dimeric, the FAS TEI is a monomer (33). All of the TEs have an α-helical insertion after strand β5 that forms a lid over the active site. Additionally, in the PKS TEIs, the N-terminal dimer-forming helices contribute to the lid structure, forming a fixed channel that runs the length of the TE and contains the active site. In contrast, the active site pocket of monomeric NRPS TEIs and TEIIs is flexible; two conformations of the lid and active site pocket were observed in the surfactin TEI (SrfTEI) crystal structure (7), and chemical shift observations suggested greater flexibility for residues of the lid region in the surfactin TEII (SrfTEII) solution structure (35). These movements seem to be of functional importance, because a movement of a linker peptide in SrfTEI determines the shape of the active site pocket and a movement of the first lid helix appears to modulate access to the active site (31).We report the structure and activity of recombinant RifR, the TEII of the rifamycin biosynthetic cluster. Steady-state kinetic analysis of the hydrolytic activity of RifR on a wide range of acyl-CoA and acyl-ACP substrates demonstrates that acyl-ACP substrates are preferred over the acyl-CoAs. Aberrant, decarboxylated acyl units are processed more efficiently than are the natural rifamycin building blocks. We report the crystal structure of RifR, the first for any hybrid PKS/NRPS TEII. The size and shape of the substrate chamber are variable, because one of the elements forming the chamber, an extended linker segment, is highly flexible, and different crystal forms reveal different shapes for the substrate binding site. Access to the active site is severely restricted, and structural comparisons with other thioesterases suggest that a conformational change in the lid and the flexible linker region is required for access to the substrate pocket.  相似文献   

17.
Members of the carotenoid cleavage dioxygenase family catalyze the oxidative cleavage of carotenoids at various chain positions, leading to the formation of a wide range of apocarotenoid signaling molecules. To explore the functions of this diverse enzyme family, we have used a chemical genetic approach to design selective inhibitors for different classes of carotenoid cleavage dioxygenase. A set of 18 arylalkyl-hydroxamic acids was synthesized in which the distance between an iron-chelating hydroxamic acid and an aromatic ring was varied; these compounds were screened as inhibitors of four different enzyme classes, either in vitro or in vivo. Potent inhibitors were found that selectively inhibited enzymes that cleave carotenoids at the 9,10 position; 50% inhibition was achieved at submicromolar concentrations. Application of certain inhibitors at 100 μm to Arabidopsis node explants or whole plants led to increased shoot branching, consistent with inhibition of 9,10-cleavage.Carotenoids are synthesized in plants and micro-organisms as photoprotective molecules and are key components in animal diets, an example being β-carotene (pro-vitamin A). The oxidative cleavage of carotenoids occurs in plants, animals, and micro-organisms and leads to the release of a range of apocarotenoids that function as signaling molecules with a diverse range of functions (1). The first gene identified as encoding a carotenoid cleavage dioxygenase (CCD)2 was the maize Vp14 gene that is required for the formation of abscisic acid (ABA), an important hormone that mediates responses to drought stress and aspects of plant development such as seed and bud dormancy (2). The VP14 enzyme cleaves at the 11,12 position (Fig. 1) of the epoxycarotenoids 9′-cis-neoxanthin and/or 9-cis-violaxanthin and is now classified as a 9-cis-epoxycarotenoid dioxygenase (NCED) (3), a subclass of the larger CCD family.Open in a separate windowFIGURE 1.Reactions catalyzed by the carotenoid cleavage dioxygenases. a, 11,12-oxidative cleavage of 9′-cis-neoxanthin by NCED; b, oxidative cleavage reactions on β-carotene and zeaxanthin.Since the discovery of Vp14, many other CCDs have been shown to be involved in the production of a variety of apocarotenoids (Fig. 1). In insects, the visual pigment retinal is formed by oxidative cleavage of β-carotene by β-carotene-15,15′-dioxygenase (4). Retinal is produced by an orthologous enzyme in vertebrates, where it is also converted to retinoic acid, a regulator of differentiation during embryogenesis (5). A distinct mammalian CCD is believed to cleave carotenoids asymmetrically at the 9,10 position (6) and, although its function is unclear, recent evidence suggests a role in the metabolism of dietary lycopene (7). The plant volatiles β-ionone and geranylacetone are produced from an enzyme that cleaves at the 9,10 position (8) and the pigment α-crocin found in the spice saffron results from an 7,8-cleavage enzyme (9).Other CCDs have been identified where biological function is unknown, for example, in cyanobacteria where a variety of cleavage specificities have been described (10-12). In other cases, there are apocarotenoids with known functions, but the identity or involvement of CCDs have not yet been described: grasshopper ketone is a defensive secretion of the flightless grasshopper Romalea microptera (13), mycorradicin is produced by plant roots during symbiosis with arbuscular mycorrhyza (14), and strigolactones (15) are plant metabolites that act as germination signals to parasitic weeds such as Striga and Orobanche (16).Recently it was discovered that strigolactones also function as a branching hormone in plants (17, 18). The existence of such a branching hormone has been known for some time, but its identity proved elusive. However, it was known that the hormone was derived from the action of at least two CCDs, max3 and max4 (more axillary growth) (19), because deletion of either of these genes in Arabidopsis thaliana, leads to a bushy phenotype (20, 21). In Escherichia coli assays, AtCCD7 (max3) cleaves β-carotene at the 9,10 position and the apocarotenoid product (10-apo-β-carotene) is reported to be further cleaved at 13,14 by AtCCD8 (max4) to produce 13-apo-β-carotene (22). Also recent evidence suggests that AtCCD8 is highly specific, cleaving only 10-apo-β-carotene (23). How the production of 13-apo-β-carotene leads to the synthesis of the complex strigolactone is unknown. The possibility remains that the enzymes may have different specificities and cleavage activities in planta. In addition, a cytochrome P450 enzyme (24) is believed to be involved in strigolactone synthesis and acts in the pathway downstream of the CCD genes. Strigolactone is thought to effect branching by regulating auxin transport (25). Because of the involvement of CCDs in strigolactone synthesis, the possibility arises that plant architecture and interaction with parasitic weeds and mycorrhyza could be controlled by the manipulation of CCD activity.Although considerable success has been obtained using genetic approaches to probe function and substrate specificity of CCDs in their native biological contexts, particularly in plant species with simple genetic systems or that are amenable to transgenesis, there are many systems where genetic approaches are difficult or impossible. Also, when recombinant CCDs are studied either in vitro or in heterologous in vivo assays, such as in E. coli strains engineered to accumulate carotenoids (26), they are often active against a broad range of substrates (5, 21, 27), and in many cases the true in vivo substrate of a particular CCD remains unknown. Therefore additional experimental tools are needed to investigate both apocarotenoid and CCD functions in their native cellular environments.In the reverse chemical genetics approach, small molecules are identified that are active against known target proteins; they are then applied to a biological system to investigate protein function in vivo (28, 29). This approach is complementary to conventional genetics since the small molecules can be applied easily to a broad range of species, their application can be controlled in dose, time, and space to provide detailed studies of biological functions, and individual proteins or whole protein classes may be targeted by varying the specificity of the small molecules. Notably, functions of the plant hormones gibberellin, brassinosteroid, and abscisic acid have been successfully probed using this approach by adapting triazoles to inhibit specific cytochrome P450 monooxygenases involved in the metabolism of these hormones (30).In the case of the CCD family, the tertiary amines abamine (31) and the more active abamineSG (32) were reported as specific inhibitors of NCED, and abamine was used to show new functions of abscisic acid in legume nodulation (33). However, no selective inhibitors for other types of CCD are known. Here we have designed a novel class of CCD inhibitor based on hydroxamic acids, where variable chain length was used to direct inhibition of CCD enzymes that cleave carotenoids at specific positions. We demonstrate the use of such novel inhibitors to control shoot branching in a model plant.  相似文献   

18.
Chromogranin A (CgA) has been proposed to play a major role in the formation of dense-core secretory granules (DCGs) in neuroendocrine cells. Here, we took advantage of unique features of the frog CgA (fCgA) to assess the role of this granin and its potential functional determinants in hormone sorting during DCG biogenesis. Expression of fCgA in the constitutively secreting COS-7 cells induced the formation of mobile vesicular structures, which contained cotransfected peptide hormones. The fCgA and the hormones coexpressed in the newly formed vesicles could be released in a regulated manner. The N- and C-terminal regions of fCgA, which exhibit remarkable sequence conservation with their mammalian counterparts were found to be essential for the formation of the mobile DCG-like structures in COS-7 cells. Expression of fCgA in the corticotrope AtT20 cells increased pro-opiomelanocortin levels in DCGs, whereas the expression of N- and C-terminal deletion mutants provoked retention of the hormone in the Golgi area. Furthermore, fCgA, but not its truncated forms, promoted pro-opiomelanocortin sorting to the regulated secretory pathway. These data demonstrate that CgA has the intrinsic capacity to induce the formation of mobile secretory granules and to promote the sorting and release of peptide hormones. The conserved terminal peptides are instrumental for these activities of CgA.Eukaryotic cells share the capacity to rapidly secrete proteins through the constitutive secretory pathway. The fundamental feature of neuroendocrine and endocrine cells is the occurrence of dense-core secretory granules (DCGs),3 which are key cytoplasmic organelles responsible for secretion of hormones, neuropeptides, and neurotransmitters through the regulated secretory pathway (RSP). Storage at high concentrations of these secretory products is required for their finely tuned release in response to extracellular stimulation (1, 2). DCG biogenesis starts with the budding of immature secretory granules (ISGs) from the trans-Golgi network (TGN) through interactions between lipid rafts and protein components, in a similar manner to constitutive vesicle budding (2, 3). The ISG budding is followed by a multistep maturation process to form the mature secretory granules, including removal of the constitutive secretory proteins and lysosomal enzymes inadvertently packaged into ISGs (4).Despite increasing knowledge of the various steps of DCG formation, the nature of the sorting signals for entry of proteins into the DCGs and the molecular machinery required to generate secretory granules are not fully elucidated (5, 6). Several recent studies highlighted the role of members of the granin family, which may represent the driving force for granulogenesis in the TGN (2), although this notion has been a matter of debate (7). Granins are soluble acidic proteins widely distributed in endocrine and neuroendocrine cells, which are characterized by the ability to aggregate at acidic pH and a high Ca2+ environment (8, 9). These conditions are found in the lumen of the TGN allowing granins to aggregate in this compartment and to be segregated from constitutively secreted proteins (10, 11). The granin aggregates are believed to associate directly or indirectly with lipid rafts at the TGN to induce budding and formation of the ISGs. A prominent role of chromogranin A (CgA) in the regulation of DCG formation in endocrine and neuroendocrine cells has been proposed. Thus, depletion of CgA in PC12 cells led to a dramatic decrease in the number of DCGs (12), and exogenously expressed CgA in these depleted PC12 cells, as in DCG-deficient endocrine A35C and 6T3 cells, restored DCG biogenesis (12, 13). Besides, expression of granins in non-endocrine, constitutively secreting cells such as CV-1, NIH3T3, or COS-7 cells provoked the formation of DCG-like structures that release their content in response to Ca2+ influx (12, 14, 15). Further investigations performed in CgA null mice and transgenic mice expressing antisense RNA against CgA also revealed a reduction in the number of DCGs in chromaffin cells that was associated with an impairment of catecholamine storage, thus demonstrating the crucial role of CgA in normal DCG biogenesis (16, 17). In CgA knockout mice, the introduction of the gene expressing human CgA restored the regulated secretory phenotype (16). A different CgA null mice strain exhibited no discernable effect on DCG formation, but elevated catecholamine secretion (18), proving that CgA deficiency is associated with hormone storage impairment in neuroendocrine cells in vivo, a finding that was confirmed in vitro (19). The CgA-/- mice strain generated by Hendy et al. (18) exhibited a compensatory overexpression of other granins, pointing to a possible overlap in granin function in secretory granule biogenesis.We reported previously that the frog CgA (fCgA) gene is coordinately regulated with the pro-opiomelanocortin (POMC) gene in the pituitary pars intermedia during the neuroendocrine reflex of skin color change, which allows amphibia to adapt to their environment through the release of POMC-derived melanotropic peptides (20, 21). Sequence comparison of fCgA with its mammalian orthologs revealed a high conservation of the N- and C-terminal domains, and far less conservation of the central part of the protein (Fig. 1A), suggesting that these domains may play a role in DCG formation and hormone release in various species (9, 20, 21). To assess the role of fCgA and its conserved N- and C-terminal regions in hormone sorting, storage, and secretion, we engineered different constructs that produce the native unmodified (no tag added) protein and truncated forms lacking the conserved N- and C-terminal domains, and we developed an antibody that specifically recognizes the central region of fCgA. Using the constitutively secreting COS-7 cells, which are devoid of DCGs, we could demonstrate for the first time that CgA is essential for targeting peptide hormones to newly formed mobile DCG-like structures. In the CgA-expressing AtT20 cells, which exhibit an only moderate capacity to sort secretory proteins to the regulated pathway (22), the granin plays a pivotal role in the sorting and release of POMC. The conserved terminal peptides of CgA are instrumental for these activities.Open in a separate windowFIGURE 1.Specificity of the antibody directed against frog CgA. A, scheme depicting the structure of fCgA and showing the high conservation of the terminal regions and the percentages of amino acid identity between frog and human CgA sequences. The highly conserved peptide WE14 and dibasic cleavage sites are also indicated. B, Western blot showing that the antibody developed against fCgA recognized the protein and several processing intermediates in frog but not rat pituitary extracts, whereas an antibody, directed against the WE14 conserved peptide, detected CgA and its processing products in both rat and frog pituitary extracts. C, immunofluorescence analysis of frog pituitary and adrenal glands, and rat adrenal gland using the antibodies against fCgA and WE14. cx, cortex; DL, distal lobe; IL, intermediate lobe; and m, medulla. Scale bars equal 10 μm.  相似文献   

19.
Viruses populate virtually every ecosystem on the planet, including the extreme acidic, thermal, and saline environments where archaeal organisms can dominate. For example, recent studies have identified crenarchaeal viruses in the hot springs of Yellowstone National Park and other high temperature environments worldwide. These viruses are often morphologically and genetically unique, with genomes that show little similarity to genes of known function, complicating efforts to understand their viral life cycles. Here, we review progress in understanding these fascinating viruses at the molecular level and the evolutionary insights coming from these studies.The last decade has seen resurgent interest in the study of viruses that lie outside traditional agricultural and medical interests. One reason is the growing appreciation of the enormous abundance and impact of viruses on the greater biosphere. For example, the oceans are thought to contain ∼1031 viruses, a truly astronomical number (1), making viruses the most abundant biological entities in this ecosystem, where they catalyze turnover of 20% of the oceanic biomass per day (1). Remarkably, the virosphere has now been shown to extend to almost every known environment on earth, including the extreme acidic, thermal, and saline environments where archaeal organisms can be dominant. Thus, because of their abundance and variety, viruses are now thought to represent the greatest reservoir of genetic diversity on the planet (2).A second reason to study archaeal viruses is a growing appreciation for the roles viruses play in evolution. Remarkably with >500 cellular genomes sequenced to date, most show a significant amount of viral or virus-like sequence within their genome, further evidence that viruses play a central role in horizontal gene transfer and help drive the evolution of their hosts. Roles for viruses in cellular evolution are also being considered. Current hypotheses contend that viruses have catalyzed several major evolutionary transitions, including the invention of DNA and DNA replication mechanisms (3), the origin of the eukaryotic nucleus (4), and thus a role in the formation of the three domains of life. In addition, there is also considerable interest in viral genesis and evolution in and of itself. To evaluate these hypotheses and to analyze evolutionary relationships among viruses, knowledge of viruses infecting the archaea is essential, yet these viruses are vastly understudied. Finally, interest in archaeal viruses stems also from the exceptional molecular insight viruses have traditionally provided into host processes; archaeal viruses are certain to provide new insights into the molecular biology of this poorly understood domain of life.Pioneering studies by Wolfram Zillig et al. (5) identified the first archaeal viruses. Although initial studies suggested that viruses infecting the euryarchaea (principally halophiles and methanogens) were similar to head-tail bacteriophage, studies of viruses infecting the hyperthermophilic crenarchaea revealed morphologies suggesting new viral families. Indeed, work by several laboratories has led to the identification of seven new viral families infecting the crenarchaea, the Globuloviridae, Guttaviridae, Fuselloviridae, Bicaudaviridae, Ampullaviridae, Rudiviridae, and Lipothrixviridae (Fig. 1) (6, 7), with STIV3 (8) and STSV1 (9) awaiting assignment. All of these viruses contain double-stranded DNA genomes ranging in size from 13.7 to 75.3 kilobase pairs, encoding 31–74 ORFs. Although many package a circular genome, the filamentous Lipothrixviridae and rod-shaped Rudiviridae are notable exceptions and are the only viruses in any domain known to encapsidate linear double-stranded DNA. Although most crenarchaeal viruses are enveloped, the Rudiviridae are devoid of lipid, and with the exception of the Fuselloviridae, they employ a lytic life cycle, although only STIV and ATV (Bicaudaviridae) are known to cause cell lysis (11).4Open in a separate windowFIGURE 1.Morphological diversity in crenarchaeal viruses. A, clockwise, beginning at upper left: STIV (8), a PSV-like virus, Sulfolobus neozealandicus droplet-shaped virus (SNDV) (47), SSV1 (48), STSV1 (9), an ATV-like virus, an SIRV virus, and S. icelandicus filamentous virus (SIFV) (10). Micrographs of SIRV, PSV-like, and ATV-like viruses from Yellowstone National Park are the courtesy of M. J. Y. Other panels are reproduced, with permission, from Refs. 810, 47, and 48. B, cryoelectron microscopy reconstruction of the STIV particle (8) showing a cutaway view (20) of the T = 31 icosahedral capsid with turret-like projections that extend from each of the 5-fold vertices. Portions of the protein shell (blue) and inner lipid layer (yellow) have been removed to reveal the interior.The exceptional morphology of these viruses has been reviewed (6, 7) and thus is only summarized here (Fig. 1). For the rod-shaped Rudiviridae, plugs are seen at both ends, from which three short tail fibers emanate, whereas the Lipothrixviridae show mop- or claw-like structures at both ends (6). Similarly, the non-tailed icosahedral viruses, STIV and euryarchaeal SH1, have large turrets or spikes that project from the surface (8, 12). In each case, these structures are thought to facilitate virus-host interactions. In contrast, other crenarchaeal viruses utilize a fusiform or lemon-shaped virion, a morphology unique to archaeal viruses. These fusiform viruses generally contain tail fibers or an extended tail on one end that is also involved in host recognition. For ATV, however, nascent particles are devoid of tails when released from the host (13). Remarkably, extended tails develop at both ends of the virion in an extracellular maturation process. Finally, Acidianus bottle-shaped virus (Ampullaviridae) shows an exceptional morphology that differs in its basic architecture from any known virus.  相似文献   

20.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  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