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 576 毫秒
1.
2.
3.
An intracellular multiplication F (IcmF) family protein is a conserved component of a newly identified type VI secretion system (T6SS) encoded in many animal and plant-associated Proteobacteria. We have previously identified ImpLM, an IcmF family protein that is required for the secretion of the T6SS substrate hemolysin-coregulated protein (Hcp) from the plant-pathogenic bacterium Agrobacterium tumefaciens. In this study, we characterized the topology of ImpLM and the importance of its nucleotide-binding Walker A motif involved in Hcp secretion from A. tumefaciens. A combination of β-lactamase-green fluorescent protein fusion and biochemical fractionation analyses revealed that ImpLM is an integral polytopic inner membrane protein comprising three transmembrane domains bordered by an N-terminal domain facing the cytoplasm and a C-terminal domain exposed to the periplasm. impLM mutants with substitutions or deletions in the Walker A motif failed to complement the impLM deletion mutant for Hcp secretion, which provided evidence that ImpLM may bind and/or hydrolyze nucleoside triphosphates to mediate T6SS machine assembly and/or substrate secretion. Protein-protein interaction and protein stability analyses indicated that there is a physical interaction between ImpLM and another essential T6SS component, ImpKL. Topology and biochemical fractionation analyses suggested that ImpKL is an integral bitopic inner membrane protein with an N-terminal domain facing the cytoplasm and a C-terminal OmpA-like domain exposed to the periplasm. Further comprehensive yeast two-hybrid assays dissecting ImpLM-ImpKL interaction domains suggested that ImpLM interacts with ImpKL via the N-terminal cytoplasmic domains of the proteins. In conclusion, ImpLM interacts with ImpKL, and its Walker A motif is required for its function in mediation of Hcp secretion from A. tumefaciens.Many pathogenic gram-negative bacteria employ protein secretion systems formed by macromolecular complexes to deliver proteins or protein-DNA complexes across the bacterial membrane. In addition to the general secretory (Sec) pathway (18, 52) and twin-arginine translocation (Tat) pathway (7, 34), which transport proteins across the inner membrane into the periplasm, at least six distinct protein secretion systems occur in gram-negative bacteria (28, 46, 66). These systems are able to secrete proteins from the cytoplasm or periplasm to the external environment or the host cell and include the well-documented type I to type V secretion systems (T1SS to T5SS) (10, 15, 23, 26, 30) and a recently discovered type VI secretion system (T6SS) (4, 8, 22, 41, 48, 49). These systems use ATPase or a proton motive force to energize assembly of the protein secretion machinery and/or substrate translocation (2, 6, 41, 44, 60).Agrobacterium tumefaciens is a soilborne pathogenic gram-negative bacterium that causes crown gall disease in a wide range of plants. Using an archetypal T4SS (9), A. tumefaciens translocates oncogenic transferred DNA and effector proteins to the host and ultimately integrates transferred DNA into the host genome. Because of its unique interkingdom DNA transfer, this bacterium has been extensively studied and used to transform foreign DNA into plants and fungi (11, 24, 40, 67). In addition to the T4SS, A. tumefaciens encodes several other secretion systems, including the Sec pathway, the Tat pathway, T1SS, T5SS, and the recently identified T6SS (72). T6SS is highly conserved and widely distributed in animal- and plant-associated Proteobacteria and plays an important role in the virulence of several human and animal pathogens (14, 19, 41, 48, 56, 63, 74). However, T6SS seems to play only a minor role or even a negative role in infection or virulence of the plant-associated pathogens or symbionts studied to date (5, 37-39, 72).T6SS was initially designated IAHP (IcmF-associated homologous protein) clusters (13). Before T6SS was documented by Pukatzki et al. in Vibrio cholerae (48), mutations in this gene cluster in the plant symbiont Rhizobium leguminosarum (5) and the fish pathogen Edwardsiella tarda (51) caused defects in protein secretion. In V. cholerae, T6SS was responsible for the loss of cytotoxicity for amoebae and for secretion of two proteins lacking a signal peptide, hemolysin-coregulated protein (Hcp) and valine-glycine repeat protein (VgrG). Secretion of Hcp is the hallmark of T6SS. Interestingly, mutation of hcp blocks the secretion of VgrG proteins (VgrG-1, VgrG-2, and VgrG-3), and, conversely, vgrG-1 and vgrG-2 are both required for secretion of the Hcp and VgrG proteins from V. cholerae (47, 48). Similarly, a requirement of Hcp for VgrG secretion and a requirement of VgrG for Hcp secretion have also been shown for E. tarda (74). Because Hcp forms a hexameric ring (41) stacked in a tube-like structure in vitro (3, 35) and VgrG has a predicted trimeric phage tail spike-like structure similar to that of the T4 phage gp5-gp27 complex (47), Hcp and VgrG have been postulated to form an extracellular translocon. This model is further supported by two recent crystallography studies showing that Hcp, VgrG, and a T4 phage gp25-like protein resembled membrane penetration tails of bacteriophages (35, 45).Little is known about the topology and structure of T6SS machinery subunits and the distinction between genes encoding machinery subunits and genes encoding regulatory proteins. Posttranslational regulation via the phosphorylation of Fha1 by a serine-threonine kinase (PpkA) is required for Hcp secretion from Pseudomonas aeruginosa (42). Genetic evidence for P. aeruginosa suggested that the T6SS may utilize a ClpV-like AAA+ ATPase to provide the energy for machinery assembly or substrate translocation (41). A recent study of V. cholerae suggested that ClpV ATPase activity is responsible for remodeling the VipA/VipB tubules which are crucial for type VI substrate secretion (6). An outer membrane lipoprotein, SciN, is an essential T6SS component for mediating Hcp secretion from enteroaggregative Escherichia coli (1). A systematic study of the T6SS machinery in E. tarda revealed that 13 of 16 genes in the evp gene cluster are essential for secretion of T6S substrates (74), which suggests the core components of the T6SS. Interestingly, most of the core components conserved in T6SS are predicted soluble proteins without recognizable signal peptide and transmembrane (TM) domains.The intracellular multiplication F (IcmF) and H (IcmH) proteins are among the few core components with obvious TM domains (8). In Legionella pneumophila Dot/Icm T4SSb, IcmF and IcmH are both membrane localized and partially required for L. pneumophila replication in macrophages (58, 70, 75). IcmF and IcmH are thought to interact with each other in stabilizing the T4SS complex in L. pneumophila (58). In T6SS, IcmF is one of the essential components required for secretion of Hcp from several animal pathogens, including V. cholerae (48), Aeromonas hydrophila (63), E. tarda (74), and P. aeruginosa (41), as well as the plant pathogens A. tumefaciens (72) and Pectobacterium atrosepticum (39). In E. tarda, IcmF (EvpO) interacted with IcmH (EvpN), EvpL, and EvpA in a yeast two-hybrid assay, and its putative nucleotide-binding site (Walker A motif) was not essential for secretion of T6SS substrates (74).In this study, we characterized the topology and interactions of the IcmF and IcmH family proteins ImpLM and ImpKL, which are two essential components of the T6SS of A. tumefaciens. We adapted the nomenclature proposed by Cascales (8), using the annotated gene designation followed by the letter indicated by Shalom et al. (59). Our data indicate that ImpLM and ImpKL are both integral inner membrane proteins and interact with each other via their N-terminal domains residing in the cytoplasm. We also provide genetic evidence showing that ImpLM may function as a nucleoside triphosphate (NTP)-binding protein or nucleoside triphosphatase to mediate T6S machinery assembly and/or substrate secretion.  相似文献   

4.
5.
6.
The extracellular chlamydial infectious particle, or elementary body (EB), is enveloped by an intra- and intermolecular cysteine cross-linked protein shell called the chlamydial outer membrane complex (COMC). A few abundant proteins, including the major outer membrane protein and cysteine-rich proteins (OmcA and OmcB), constitute the overwhelming majority of COMC proteins. The identification of less-abundant COMC proteins has been complicated by limitations of proteomic methodologies and the contamination of COMC fractions with abundant EB proteins. Here, we used parallel liquid chromatography-mass spectrometry/mass spectrometry (LC-MS/MS) analyses of Chlamydia trachomatis serovar L2 434/Bu EB, COMC, and Sarkosyl-soluble EB fractions to identify proteins enriched or depleted from COMC. All well-described COMC proteins were specifically enriched in the COMC fraction. In contrast, multiple COMC-associated proteins found in previous studies were strongly enriched in the Sarkosyl-soluble fraction, suggesting that these proteins are not COMC components or are not stably associated with COMC. Importantly, we also identified novel proteins enriched in COMC. The list of COMC proteins identified in this study has provided reliable information for further understanding chlamydial protein secretion systems and modeling COMC and EB structures.Bacteria in the phylum Chlamydiae are characterized by their complex intracellular developmental cycles. Chlamydiae must assume at least two functionally distinct morphotypes, the intracellular, replicative reticulate body (RB) and the extracellular, infectious elementary body (EB), to replicate and be transmitted to new hosts (50). The divergence of distinct RB and EB forms may have been driven by the different pressures these pathogens face inside host cells during replication and outside host cells during transmission. For example, the outer membrane of EB contains a poorly immunogenic truncated lipopolysaccharide (LPS) (14, 30) and immunodominant epitopes of the major outer membrane protein (MOMP) vary substantially among closely related chlamydial strains (13). EB also lack detectable peptidoglycan (2, 20, 60), although functional murein biosynthetic enzymes (2, 5, 16, 21, 32, 43, 45, 46) are expressed in RB during productive and persistent infection (44). To compensate for the loss of murein, EB are enveloped by a protein P-layer, which lends osmotic stability to the infectious particle (29).Attempts to identify components of the P-layer and outer membrane proteins of Chlamydia were advanced by the observation that these layers can be separated from many soluble EB proteins using the detergent N-lauroyl sarcosine (Sarkosyl). Caldwell et al. dubbed the Sarkosyl-insoluble fraction the chlamydial outer membrane complex (COMC) and noted that purified COMC maintained the shape of intact EB and contained a complete outer membrane, and they reported that a single outer membrane protein, MOMP, accounted for more than 60% of total COMC protein content (15). Other studies revealed that the COMC is stabilized by extensive disulfide bonds between MOMP monomers (26, 27, 53) and between MOMP and two abundant cysteine-rich COMC components (26, 28). Other studies revealed that the COMC is stabilized by extensive disulfide bonds between MOMP monomers (18, 29, 52) and the EB surface (3, 19, 47, 67). More recent data suggest that not all EB outer membrane (OM) proteins are disulfide cross-linked to the COMC. For example, polymorphic membrane protein D localizes to the surface of EB but can be extracted from intact EB with gentle detergents in the absence of reducing agents (17, 61). Thus, not all COMC proteins are exposed on the EB surface, nor are all EB OM proteins components of the COMC.Beyond these well-described and abundant COMC components, other studies have indicated that additional proteins localize to the EB surface and/or COMC of Chlamydia trachomatis (7, 28, 36, 51, 57, 64, 67, 70). However, confirming that specific proteins localize to the COMC or OM of EB can be challenging due to factors such as the contamination of EB preparations with RB proteins and technical limitations of proteomic and surface-labeling protein identification methods (29, 56).Here, we used differential proteomics to identify proteins specifically enriched in the COMC. Isolated COMC were dissolved in 8 M urea, and the extracted proteins were digested with trypsin. The resulting peptides were analyzed by high-sensitivity liquid chromatography-mass spectrometry/mass spectrometry (LC-MS/MS) to identify low-abundance proteins. Sarkosyl-soluble fractions and whole EB were analyzed in parallel with COMC, and protein assignments were compared among three replicate runs of each fraction. In total, peptides from 329 L2 proteins were identified. The differential analysis of protein abundance indicated the enrichment of 17 proteins in the COMC. Our results define the cadre of low-abundance COMC proteins, provide a starting point for the identification of surface-exposed EB proteins, and identify EB proteins that are likely to be recognized by innate immunity receptors and/or capable of eliciting neutralizing antibodies in vivo. Finally, our findings and data from other recent studies permit the refinement of existing models of EB and COMC structure.  相似文献   

7.
The sequestration of iron by mammalian hosts represents a significant obstacle to the establishment of a bacterial infection. In response, pathogenic bacteria have evolved mechanisms to acquire iron from host heme. Bacillus anthracis, the causative agent of anthrax, utilizes secreted hemophores to scavenge heme from host hemoglobin, thereby facilitating iron acquisition from extracellular heme pools and delivery to iron-regulated surface determinant (Isd) proteins covalently attached to the cell wall. However, several Gram-positive pathogens, including B. anthracis, contain genes that encode near iron transporter (NEAT) proteins that are genomically distant from the genetically linked Isd locus. NEAT domains are protein modules that partake in several functions related to heme transport, including binding heme and hemoglobin. This finding raises interesting questions concerning the relative role of these NEAT proteins, relative to hemophores and the Isd system, in iron uptake. Here, we present evidence that a B. anthracis S-layer homology (SLH) protein harboring a NEAT domain binds and directionally transfers heme to the Isd system via the cell wall protein IsdC. This finding suggests that the Isd system can receive heme from multiple inputs and may reflect an adaptation of B. anthracis to changing iron reservoirs during an infection. Understanding the mechanism of heme uptake in pathogenic bacteria is important for the development of novel therapeutics to prevent and treat bacterial infections.Pathogenic bacteria need to acquire iron to survive in mammalian hosts (12). However, the host sequesters most iron in the porphyrin heme, and heme itself is often bound to proteins such as hemoglobin (14, 28, 85). Circulating hemoglobin can serve as a source of heme-iron for replicating bacteria in infected hosts, but the precise mechanisms of heme extraction, transport, and assimilation remain unclear (25, 46, 79, 86). An understanding of how bacterial pathogens import heme will lead to the development of new anti-infectives that inhibit heme uptake, thereby preventing or treating infections caused by these bacteria (47, 68).The mechanisms of transport of biological molecules into a bacterial cell are influenced by the compositional, structural, and topological makeup of the cell envelope. Gram-negative bacteria utilize specific proteins to transport heme through the outer membrane, periplasm, and inner membrane (83, 84). Instead of an outer membrane and periplasm, Gram-positive bacteria contain a thick cell wall (59, 60). Proteins covalently anchored to the cell wall provide a functional link between extracellular heme reservoirs and intracellular iron utilization pathways (46). In addition, several Gram-positive and Gram-negative bacterial genera also contain an outermost structure termed the S (surface)-layer (75). The S-layer is a crystalline array of protein that surrounds the bacterial cell and may serve a multitude of functions, including maintenance of cell architecture and protection from host immune components (6, 7, 18, 19, 56). In bacterial pathogens that manifest an S-layer, the “force field” function of this structure raises questions concerning how small molecules such as heme can be successfully passed from the extracellular milieu to cell wall proteins for delivery into the cell cytoplasm.Bacillus anthracis is a Gram-positive, spore-forming bacterium that is the etiological agent of anthrax disease (30, 33). The life cycle of B. anthracis begins after a phagocytosed spore germinates into a vegetative cell inside a mammalian host (2, 40, 69, 78). Virulence determinants produced by the vegetative cells facilitate bacterial growth, dissemination to major organ systems, and eventually host death (76-78). The release of aerosolized spores into areas with large concentrations of people is a serious public health concern (30).Heme acquisition in B. anthracis is mediated by the action of IsdX1 and IsdX2, two extracellular hemophores that extract heme from host hemoglobin and deliver the iron-porphyrin to cell wall-localized IsdC (21, 45). Both IsdX1 and IsdX2 harbor near iron transporter domains (NEATs), a conserved protein module found in Gram-positive bacteria that mediates heme uptake from hemoglobin and contributes to bacterial pathogenesis upon infection (3, 8, 21, 31, 44, 46, 49, 50, 67, 81, 86). Hypothesizing that B. anthracis may contain additional mechanisms for heme transport, we provide evidence that B. anthracis S-layer protein K (BslK), an S-layer homology (SLH) and NEAT protein (32, 43), is surface localized and binds and transfers heme to IsdC in a rapid, contact-dependent manner. These results suggest that the Isd system is not a self-contained conduit for heme trafficking and imply that there is functional cross talk between differentially localized NEAT proteins to promote heme uptake during infection.  相似文献   

8.
The respiratory syncytial virus (RSV) matrix (M) protein is localized in the nucleus of infected cells early in infection but is mostly cytoplasmic late in infection. We have previously shown that M localizes in the nucleus through the action of the importin β1 nuclear import receptor. Here, we establish for the first time that M''s ability to shuttle to the cytoplasm is due to the action of the nuclear export receptor Crm1, as shown in infected cells, and in cells transfected to express green fluorescent protein (GFP)-M fusion proteins. Specific inhibition of Crm1-mediated nuclear export by leptomycin B increased M nuclear accumulation. Analysis of truncated and point-mutated M derivatives indicated that Crm1-dependent nuclear export of M is attributable to a nuclear export signal (NES) within residues 194 to 206. Importantly, inhibition of M nuclear export resulted in reduced virus production, and a recombinant RSV carrying a mutated NES could not be rescued by reverse genetics. That this is likely to be due to the inability of a nuclear export deficient M to localize to regions of virus assembly is indicated by the fact that a nuclear-export-deficient GFP-M fails to localize to regions of virus assembly when expressed in cells infected with wild-type RSV. Together, our data suggest that Crm1-dependent nuclear export of M is central to RSV infection, representing the first report of such a mechanism for a paramyxovirus M protein and with important implications for related paramyxoviruses.The Pneumovirus respiratory syncytial virus (RSV) within the Paramyxoviridae family is the most common cause of lower-respiratory-tract disease in infants (7). The negative-sense single-strand RNA genome of RSV encodes two nonstructural and nine structural proteins, comprising the envelope glycoproteins (F, G, and SH), the nucleocapsid proteins (N, P, and L), the nucleocapsid-associated proteins (M2-1 and M2-2), and the matrix (M) protein (1, 7, 11). Previously, we have shown that M protein localizes in the nucleus at early stages of infection, but later in infection it is localized mainly in the cytoplasm, in association with nucleocapsid-containing cytoplasmic inclusions (13, 16). The M proteins of other negative-strand viruses, such as Sendai virus, Newcastle disease virus, and vesicular stomatitis virus (VSV), have also been observed in the nucleus at early stages of infection (32, 40, 48). Interestingly, the M proteins of all of these viruses, including RSV, play major roles in virus assembly, which take place in the cytoplasm and at the cell membrane (11, 12, 24, 34, 36, 39), but the mechanisms by which trafficking between the nucleus and cytoplasm occurs are unknown.The importin β family member Crm1 (exportin 1) is known to mediate nuclear export of proteins bearing leucine-rich nuclear export signals (NES) (8, 9, 18, 19, 37, 42, 43), such as the human immunodeficiency virus type 1 Rev protein (4). In the case of the influenza virus matrix (M1) protein, binding to the influenza virus nuclear export protein, which possesses a Crm1-recognized NES, appears to be responsible for its export from the nucleus, bound to the influenza virus RNA (3).We have recently shown that RSV M localizes in the nucleus through a conventional nuclear import pathway dependent on the nuclear import receptor importin β1 (IMPβ1) and the guanine nucleotide-binding protein Ran (14). In the present study, we show for the first time that RSV M possesses a Crm1-dependent nuclear export pathway, based on experiments using the specific inhibitor leptomycin B (LMB) (25), both in RSV-infected cells and in green fluorescent protein (GFP)-M fusion protein-expressing transfected cells. We use truncated and point-mutated M derivatives to map the Crm1-recognized NES within the M sequence and show that Crm1-dependent nuclear export is critical to the RSV infectious cycle, since LMB treatment early in infection, inhibiting M export from the nucleus, reduces RSV virion production and a recombinant RSV carrying a NES mutation in M was unable to replicate, probably because M deficient in nuclear export could not localize to areas of virus assembly, as shown in RSV-infected cells transfected to express GFP-M. This is the first report of a Crm1-mediated nuclear export pathway for a paramyxovirus M protein, with implications for the trafficking and function of other paramyxovirus M proteins.  相似文献   

9.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

10.
11.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

12.
13.
14.
A bioinformatic analysis of nearly 400 genomes indicates that the overwhelming majority of bacteria possess homologs of the Escherichia coli proteins FtsL, FtsB, and FtsQ, three proteins essential for cell division in that bacterium. These three bitopic membrane proteins form a subcomplex in vivo, independent of the other cell division proteins. Here we analyze the domains of E. coli FtsL that are involved in the interaction with other cell division proteins and important for the assembly of the divisome. We show that FtsL, as we have found previously with FtsB, packs an enormous amount of information in its sequence for interactions with proteins upstream and downstream in the assembly pathway. Given their size, it is likely that the sole function of the complex of these two proteins is to act as a scaffold for divisome assembly.The division of an Escherichia coli cell into two daughter cells requires a complex of proteins, the divisome, to coordinate the constriction of the three layers of the Gram-negative cell envelope. In E. coli, there are 10 proteins known to be essential for cell division; in the absence of any one of these proteins, cells continue to elongate and to replicate and segregate their chromosomes but fail to divide (29). Numerous additional nonessential proteins have been identified that localize to midcell and assist in cell division (7-9, 20, 25, 34, 56, 59).A localization dependency pathway has been determined for the 10 essential division proteins (FtsZ→FtsA/ZipA→FtsK→FtsQ→FtsL/FtsB→FtsW→FtsI→FtsN), suggesting that the divisome assembles in a hierarchical manner (29). Based on this pathway, a given protein depends on the presence of all upstream proteins (to the left) for its localization and that protein is then required for the localization of the downstream division proteins (to the right). While the localization dependency pathway of cell division proteins suggests that a sequence of interactions is necessary for divisome formation, recent work using a variety of techniques reveals that a more complex web of interactions among these proteins is necessary for a functionally stable complex (6, 10, 19, 23, 24, 30-32, 40). While numerous interactions have been identified between division proteins, further work is needed to define which domains are involved and which interactions are necessary for assembly of the divisome.One subcomplex of the divisome, composed of the bitopic membrane proteins FtsB, FtsL, and FtsQ, appears to be the bridge between the predominantly cytoplasmic cell division proteins and the predominantly periplasmic cell division proteins (10). FtsB, FtsL, and FtsQ share a similar topology: short amino-terminal cytoplasmic domains and larger carboxy-terminal periplasmic domains. This tripartite complex can be divided further into a subcomplex of FtsB and FtsL, which forms in the absence of FtsQ and interacts with the downstream division proteins FtsW and FtsI in the absence of FtsQ (30). The presence of an FtsB/FtsL/FtsQ subcomplex appears to be evolutionarily conserved, as there is evidence that the homologs of FtsB, FtsL, and FtsQ in the Gram-positive bacteria Bacillus subtilis and Streptococcus pneumoniae also assemble into complexes (18, 52, 55).The assembly of the FtsB/FtsL/FtsQ complex is important for the stabilization and localization of one or more of its component proteins in both E. coli and B. subtilis (11, 16, 18, 33). In E. coli, FtsB and FtsL are codependent for their stabilization and for localization to midcell, while FtsQ does not require either FtsB or FtsL for its stabilization or localization to midcell (11, 33). Both FtsL and FtsB require FtsQ for localization to midcell, and in the absence of FtsQ the levels of full-length FtsB are significantly reduced (11, 33). The observed reduction in full-length FtsB levels that occurs in the absence of FtsQ or FtsL results from the degradation of the FtsB C terminus (33). However, the C-terminally degraded FtsB generated upon depletion of FtsQ can still interact with and stabilize FtsL (33).While a portion of the FtsB C terminus is dispensable for interaction with FtsL and for the recruitment of the downstream division proteins FtsW and FtsI, it is required for interaction with FtsQ (33). Correspondingly, the FtsQ C terminus also appears to be important for interaction with FtsB and FtsL (32, 61). The interaction between FtsB and FtsL appears to be mediated by the predicted coiled-coil motifs within the periplasmic domains of the two proteins, although only the membrane-proximal half of the FtsB coiled coil is necessary for interaction with FtsL (10, 32, 33). Additionally, the transmembrane domains of FtsB and FtsL are important for their interaction with each other, while the cytoplasmic domain of FtsL is not necessary for interaction with FtsB, which has only a short 3-amino-acid cytoplasmic domain (10, 33).In this study, we focused on the interaction domains of FtsL. We find that, as with FtsB, the C terminus of FtsL is required for the interaction of FtsQ with the FtsB/FtsL subcomplex while the cytoplasmic domain of FtsL is involved in recruitment of the downstream division proteins. Finally, we provide a comprehensive analysis of the presence of FtsB, FtsL, and FtsQ homologs among bacteria and find that the proteins of this complex are likely more widely distributed among bacteria than was previously thought.  相似文献   

15.
Agrobacterium tumefaciens VirB proteins assemble a type IV secretion apparatus and a T-pilus for secretion of DNA and proteins into plant cells. The pilin-like protein VirB3, a membrane protein of unknown topology, is required for the assembly of the T-pilus and for T-DNA secretion. Using PhoA and green fluorescent protein (GFP) as periplasmic and cytoplasmic reporters, respectively, we demonstrate that VirB3 contains two membrane-spanning domains and that both the N and C termini of the protein reside in the cytoplasm. Fusion proteins with GFP at the N or C terminus of VirB3 were fluorescent and, like VirB3, localized to a cell pole. Biochemical fractionation studies demonstrated that VirB3 proteins encoded by three Ti plasmids, the octopine Ti plasmid pTiA6NC, the supervirulent plasmid pTiBo542, and the nopaline Ti plasmid pTiC58, are inner membrane proteins and that VirB4 has no effect on membrane localization of pTiA6NC-encoded VirB3 (pTiA6NC VirB3). The pTiA6NC and pTiBo542 VirB2 pilins, like VirB3, localized to the inner membrane. The pTiC58 VirB4 protein was earlier found to be essential for stabilization of VirB3. Stabilization of pTiA6NC VirB3 requires not only VirB4 but also two additional VirB proteins, VirB7 and VirB8. A binary interaction between VirB3 and VirB4/VirB7/VirB8 is not sufficient for VirB3 stabilization. We hypothesize that bacteria use selective proteolysis as a mechanism to prevent assembly of unproductive precursor complexes under conditions that do not favor assembly of large macromolecular structures.Bacteria use type IV secretion (T4S) to deliver macromolecules to prokaryotes and eukaryotes (12). Animal and human pathogens deliver proteins to their eukaryotic hosts to affect cellular processes causing disease. The plant-pathogenic bacterium Agrobacterium tumefaciens delivers both proteins and DNA to plants and other eukaryotes. DNA delivered by Agrobacterium directs constitutive synthesis of phytohormones in a transformed plant cell, promoting cancerous growth (56). The Ptl toxin of Bordetella pertussis modifies G proteins by ADP-ribosylation, affecting intracellular cell signaling, and CagA of Helicobacter pylori disrupts epithelial cell polarity by inhibiting PAR1 kinase activity (37, 44, 47). T4S is ancestrally related to bacterial conjugation, a mechanism used by bacteria for interbacterial plasmid transfer, enabling them to acquire novel genes for antibiotic resistance, degradation of organic molecules, toxin production, and other virulence traits (29).The VirD4/VirB family of proteins, found conserved in many alphaproteobacteria, mediates T4S (12). The Ti plasmid-encoded Agrobacterium T4S system requires VirD4 and 11 VirB proteins, VirB1 to VirB11, for efficient DNA transfer (7, 54). The membrane and membrane-associated VirB proteins assemble a macromolecular structure at the cell membrane to promote substrate transfer (12). The octopine Ti plasmid pTiA6NC-encoded VirB6 to VirB11 proteins assemble the T4S apparatus at a cell pole (34, 35, 39). The VirD4 coupling protein targets the VirE2 substrate protein to the cell pole (4). A recent study found that the nopaline Ti plasmid pTiC58 T4S system (T4SS) and its substrates form a helical array around the cell circumference (1). Structural studies using Escherichia coli conjugative plasmid pKM101-encoded VirB homologues showed that TraN (VirB7), TraO (VirB9), and TraF (VirB10) form the core complex and that TraF forms a channel at the outer membrane (11, 23). The Agrobacterium VirB proteins assemble a T-pilus, an appendage composed primarily of VirB2, with VirB5 and VirB7 as its minor constituents (38, 40, 41, 48, 50, 55). VirB3, a homolog of the pilin-like TraL protein encoded in E. coli plasmids, is postulated to function in T-pilus assembly (52). Three ATP-utilizing proteins, VirB4, VirB11, and VirD4, supply energy for substrate translocation (3, 9, 34).The membrane topology of all the VirB proteins, except for VirB3, was determined by analyses of random phoA insertion mutants, targeted phoA fusions, and targeted bla fusions (6, 14, 15, 21, 22, 31, 35, 53). phoA and bla, which encode alkaline phosphatase and β-lactamase, respectively, serve as excellent markers for periplasmic proteins, as they are enzymatically active only when targeted to the cell periplasm (8, 30). Green fluorescent protein (GFP) is an ideal cytoplasmic marker because it fluoresces only when located in the cytoplasm (19, 20). When GFP is targeted to the periplasm through fusion with a membrane-spanning domain (MSD), it fails to fold properly and does not fluoresce.The prevailing view, based on in silico analysis, is that VirB3 is a bitopic membrane protein with a periplasmic C terminus. No phoA-positive insertions in virB3, however, were identified in two random mutagenesis studies of the virB operon (6, 15). The small size of VirB3, a polypeptide of 108 amino acids (aa), could be a contributing factor to the negative findings. Yet several PhoA-positive insertions in two smaller VirB proteins, VirB2 (74-aa mature peptide) and VirB7 (41-aa mature peptide), were successfully obtained in both studies. Therefore, the negative findings may also be indicative of the presence of a small periplasmic domain in VirB3. Biochemical studies showed that the nopaline Ti plasmid pTiC58-encoded VirB3 protein (pTiC58 VirB3) associates with the bacterial outer membrane, while VirB2 associates with both the inner and outer membranes (52). The pTiC58 VirB4 protein is required for localization of VirB3 to the outer membrane (33). VirB4 is also required for VirB3 stability (33, 55). A low level of VirB3 accumulated in a nonpolar pTiC58 virB6 deletion mutant; however, addition of virB6 in trans did not restore the level of the protein, even though it restored tumorigenicity (27). VirB3 participates in the formation of protein complexes with the T-pilus proteins VirB2 and VirB5 (55).Homologues of VirB3 are found in many alphaproteobacteria with a T4SS. While most VirB3 homologues are small proteins, several recently identified homologues are fusions of VirB3 and the immediate downstream protein VirB4 (5, 10, 24). These fusion homologs, which include Actinobacillus MagB03 (GenBank accession no. AAG24434), Campylobacter CmgB3/4 (EAQ71805), Yersinia pseudotuberculosis TriC (CAF25448), Citrobacter koseri PilX3-4 (ABV12046), and Klebsiella pneumoniae PilX3-4 (BAF49490), have VirB3 at the N terminus and VirB4 at the C terminus. Agrobacterium VirB4 is an integral membrane protein with a cytoplasmic N terminus (14). Its homologues are expected to have a similar topology. The prevailing view that pTi VirB3 has a periplasmic C terminus is inconsistent with the cytoplasmic location of the N terminus of VirB4 in the VirB3-VirB4 fusion protein homologues.In this study, we report the membrane topology of Agrobacterium VirB3 and demonstrate that the C terminus of the protein resides in the cytoplasm. We also demonstrate that VirB3 is an inner membrane protein, not an outer membrane protein as previously reported (52). The octopine Ti plasmid pTiA6NC VirB4 protein does not affect membrane localization of VirB3 but does stabilize VirB3. VirB4, however, is not sufficient for pTiA6NC VirB3 stabilization. Two additional proteins, VirB7 and VirB8, are required for the stabilization of pTiA6NC VirB3.  相似文献   

16.
17.
Secretion of cholera toxin and other virulence factors from Vibrio cholerae is mediated by the type II secretion (T2S) apparatus, a multiprotein complex composed of both inner and outer membrane proteins. To better understand the mechanism by which the T2S complex coordinates translocation of its substrates, we are examining the protein-protein interactions of its components, encoded by the extracellular protein secretion (eps) genes. In this study, we took a cell biological approach, observing the dynamics of fluorescently tagged EpsC and EpsM proteins in vivo. We report that the level and context of fluorescent protein fusion expression can have a bold effect on subcellular location and that chromosomal, intraoperon expression conditions are optimal for determining the intracellular locations of fusion proteins. Fluorescently tagged, chromosomally expressed EpsC and EpsM form discrete foci along the lengths of the cells, different from the polar localization for green fluorescent protein (GFP)-EpsM previously described, as the fusions are balanced with all their interacting partner proteins within the T2S complex. Additionally, we observed that fluorescent foci in both chromosomal GFP-EpsC- and GFP-EpsM-expressing strains disperse upon deletion of epsD, suggesting that EpsD is critical to the localization of EpsC and EpsM and perhaps their assembly into the T2S complex.The type II secretion (T2S) pathway is widely used by pathogenic gram-negative bacteria for delivery of virulence factors into the extracellular milieu (11, 17, 46). Proteins destined for release through this pathway are first translocated across the cytoplasmic membrane via the Sec (24, 42) or Tat (59) machinery. Following folding and assembly in the periplasm, the proteins are transported across the outer membrane via the T2S machinery, a complex composed of 12 to 16 different gene products, depending on the species. In Vibrio cholerae, the elements of the T2S apparatus are encoded by the extracellular protein secretion (eps) genes, epsC through epsN and pilD (vcpD) (18, 31, 39, 49, 50). Together these proteins coordinate the outer membrane translocation of the major virulence factor, cholera toxin, as well as chitinase, lipase, hemagglutinin/protease, and other proteases (12, 27, 49). Our studies are focused on better understanding how the T2S complex assembles in the cell envelope of V. cholerae to begin to elucidate the mechanism by which extracellular secretion is accomplished.The T2S apparatus is modeled as an envelope-spanning complex with subcomplexes in the inner and outer membranes (see Fig. S1 in the supplemental material). The precise stoichiometry and juxtaposition of the Eps proteins are not known, but accumulating biochemical, genetic, and molecular studies continue to refine our understanding of complex assembly and function (for a review, see reference 25). A trimolecular complex consisting of cytoplasmic protein EpsE and inner membrane proteins EpsL and EpsM has been identified. EpsL and EpsM have been shown to coimmunoprecipitate and participate in mutual stabilization interactions in vivo by protecting each other from proteolysis (34, 41, 43, 48). Homologs of inner membrane protein EpsC have been implicated in interactions with the aforementioned inner membrane subcomplex (20, 29, 57), as well as homologs of outer membrane protein EpsD, which form oligomeric rings through which the secreted substrates, it is hypothesized, exit the cell (1, 10, 36, 38). More specifically, EpsC homologs in Pseudomonas aeruginosa and Klebsiella oxytoca are sensitive to proteolysis or unable to oligomerize in the absence of EpsD homologs (2, 40); however, direct interactions between these two proteins in their full-length forms have not been shown by coimmunoprecipitation or copurification. Although yeast two-hybrid analysis of the periplasmic domains of the Erwinia chrysanthemi EpsC and EpsD homologs also did not reveal interaction (15), recently it was shown that periplasmic subdomains of EpsC and EpsD homologs of Vibrio vulnificus copurified (28). It seems likely that EpsC, having interactions with both inner and outer membrane subcomplexes, plays a crucial role in complex function by connecting the inner membrane components to the outer membrane EpsD pore. Furthermore, it has been speculated that EpsC homologs impart specificity to the various T2S systems by directly interacting with proteins to be secreted (3).We have taken a cell biology approach to characterizing Eps protein interactions, observing the dynamics of green fluorescent protein (GFP)-tagged components of the Eps complex in live cells by fluorescence microscopy. This method permits study of Eps protein assembly in the context of the complete apparatus, situated in both membranes, without the disruptive procedures required for many in vitro molecular and biochemical analyses of protein-protein interactions. Here we present data illustrating the importance of expressing GFP fusions for localization studies with all other interacting components, preserving wild-type stoichiometry and expression levels. In particular, we note that GFP-EpsM does not appear to be focused at the polar membrane as previously described (53), when expressed in balance with its interacting proteins. Chromosomal replacement of epsM and epsC with gfp-tagged versions instead reveals a more distributed pattern, with punctate fluorescent foci along the full length of the cell. We have exploited these chromosomal gfp-eps strains to further dissect the interactions and requirements for localization of EpsC and EpsM by systematically deleting other eps genes in the operon.  相似文献   

18.
19.
For construction of the bacterial flagellum, many of the flagellar proteins are exported into the central channel of the flagellar structure by the flagellar type III protein export apparatus. FlhA and FlhB, which are integral membrane proteins of the export apparatus, form a docking platform for the soluble components of the export apparatus, FliH, FliI, and FliJ. The C-terminal cytoplasmic domain of FlhA (FlhAC) is required for protein export, but it is not clear how it works. Here, we analyzed a temperature-sensitive Salmonella enterica mutant, the flhA(G368C) mutant, which has a mutation in the sequence encoding FlhAC. The G368C mutation did not eliminate the interactions with FliH, FliI, FliJ, and the C-terminal cytoplasmic domain of FlhB, suggesting that the mutation blocks the export process after the FliH-FliI-FliJ-export substrate complex binds to the FlhA-FlhB platform. Limited proteolysis showed that FlhAC consists of at least three subdomains, a flexible linker, FlhACN, and FlhACC, and that FlhACN becomes sensitive to proteolysis by the G368C mutation. Intragenic suppressor mutations were identified in these subdomains and restored flagellar protein export to a considerable degree. However, none of these suppressor mutations suppressed the protease sensitivity. We suggest that FlhAC not only forms part of the docking platform for the FliH-FliI-FliJ-export substrate complex but also is directly involved in the translocation of the export substrate into the central channel of the growing flagellar structure.The bacterial flagellum, which is responsible for motility, is a supramolecular complex of about 30 different proteins, and it consists of at least three substructures: the basal body, the hook, and the filament. Flagellar assembly begins with the basal body, followed by the hook and finally the filament. Many of the flagellar component proteins are translocated into the central channel of the growing flagellar structure and then to the distal end of the structure for self-assembly by the flagellar type III protein export apparatus (11, 16, 22). This export apparatus consists of six integral membrane proteins, FlhA, FlhB, FliO, FliP, FliQ, and FliR, and three soluble proteins, FliH, FliI, and FliJ (18, 21). These protein components show significant sequence and functional similarities to those of the type III secretion systems of pathogenic bacteria, which directly inject virulence factors into their host cells (11, 16).FliI is an ATPase (4) and forms an FliH2-FliI complex with its regulator, FliH, in the cytoplasm (20). FliI self-assembles into a homo-hexamer and hence exhibits full ATPase activity (1, 8, 17). FliH and FliI, together with FliJ and the export substrate, bind to the export core complex, which is composed of the six integral membrane proteins, to recruit export substrates from the cytoplasm to the core complex (14) and facilitate the initial entry of export substrates into the export gate (23). FliJ not only prevents premature aggregation of export substrates in the cytoplasm (13) but also plays an important role in the escort mechanism for cycling export chaperones during flagellar assembly (3). The export core complex is believed to be located in the central pore of the basal body MS ring (11, 16, 22). In fact, it has been found that FlhA, FliP, and FliR are associated with the MS ring (5, 9). The FliR-FlhB fusion protein is partially functional, suggesting that FliR and FlhB interact with each other within the MS ring (29). The export core complex utilizes a proton motive force across the cytoplasmic membrane as the energy source to drive the successive unfolding of export substrates and their translocation into the central channel of the growing flagellum (23, 27). Here we refer to the export core complex as the “export gate,” as we have previously (8, 16, 23, 24).FlhA is a 692-amino-acid protein consisting of two regions: a hydrophobic N-terminal transmembrane region with eight predicted α-helical transmembrane spans (FlhATM) and a hydrophilic C-terminal cytoplasmic region (FlhAC) (12, 15). FlhATM is responsible for the association with the MS ring (9). FlhAC interacts with FliH, FliI, FliJ, and the C-terminal cytoplasmic domain of FlhB (6, 12, 21, 24) and plays a role in the initial export process with these proteins (28). It has been shown that the V404M mutation in FlhAC increases not only the probability of FliI binding to the export gate in the absence of FliH (14) but also the efficiency of substrate translocation through the export gate in the absence of FliH and FliI (23). Recently, it has been shown that FlhAC is also required for substrate recognition (7). These observations suggest that an interaction between FlhAC and FliI is coupled with substrate entry, although it is not clear how.In order to understand the mechanism of substrate entry into the export gate, we characterized a temperature-sensitive Salmonella enterica mutant, the flhA(G368C) mutant, whose mutation blocks the flagellar protein export process at 42°C (28). We show here that this mutation severely inhibits translocation of flagellar proteins through the export gate after the FliH-FliI-FliJ complex binds to the FlhA-FlhB platform of the gate and that the impaired ability of the flhA(G368C) mutant to export flagellar proteins is restored almost to wild-type levels by intragenic second-site mutations that may alter the interactions between subdomains of FlhAC for possible rearrangement for the export function.  相似文献   

20.
Glutathione constitutes a key player in the thiol redox buffer in many organisms. However, the gram-positive bacteria Bacillus subtilis and Staphylococcus aureus lack this low-molecular-weight thiol. Recently, we identified S-cysteinylated proteins in B. subtilis after treatment of cells with the disulfide-generating electrophile diamide. S cysteinylation is thought to protect protein thiols against irreversible oxidation to sulfinic and sulfonic acids. Here we show that S thiolation occurs also in S. aureus proteins after exposure to diamide. We further analyzed the formation of inter- and intramolecular disulfide bonds in cytoplasmic proteins using diagonal nonreducing/reducing sodium dodecyl sulfate gel electrophoresis. However, only a few proteins were identified that form inter- or intramolecular disulfide bonds under control and diamide stress conditions in B. subtilis and S. aureus. Depletion of the cysteine pool was concomitantly measured in B. subtilis using a metabolomics approach. Thus, the majority of reversible thiol modifications that were previously detected by two-dimensional gel fluorescence-based thiol modification assay are most likely based on S thiolations. Finally, we found that a glutathione-producing B. subtilis strain which expresses the Listeria monocytogenes gshF gene did not show enhanced oxidative stress resistance compared to the wild type.Cysteine thiols in proteins fulfill an important and diverse set of cellular functions. In particular, they participate in enzymatic catalysis; in metal coordination, such as in the generation of Fe-S-clusters; and in determining the spatial structure of proteins via disulfide bond formation (3, 22, 23, 38). Cysteines are strong nucleophiles amenable to posttranslational modifications by reactive oxygen species (ROS) and reactive nitrogen species, leading to disulfides; to sulfenic, sulfinic, or sulfonic acids; mixed disulfides with low-molecular-weight (LMW) thiols (S thiolations); and S nitrosylations (7, 16, 17, 27).The redox status of the cytoplasm is under physiological conditions in a reduced state. Thus, most cysteines are present as free thiols (6). Because aerobic organisms have to cope with oxidative stress caused by ROS, such as superoxide anions, hydrogen peroxide, or hydroxyl radicals, they need to employ effective mechanisms that maintain the reduced state. In gram-negative bacteria, the thiol-disulfide balance is accomplished by the glutathione (GSH) system, a thiol-based redox buffer. The GSH system consists of glutaredoxin (Grx), GSH (γ-glutamylcysteinyl glycine), GSH reductase, and GSH peroxidase (34). Reduction of disulfides occurs via sequential electron transfer from glutaredoxin and reduced GSH; oxidized GSH (GSSG) is reduced by the NADPH-dependent GSH reductase. GSH peroxidase enables the direct detoxification of ROS by GSH oxidation.However, many gram-positive bacteria lack genes for GSH biosynthesis. Actinomycetes instead use a thiol redox buffer based on mycothiol (50). Bacillus subtilis, Staphylococcus aureus, and other gram-positive bacteria rely on different thiol redox buffers based on cysteine, the novel 398-Da bacillithiol (BSH), or coenzyme A (CoA) (15, 52). To maintain the reduced state of the cytoplasm, most bacteria use enzymatic systems for disulfide bond reduction such as the thioredoxin (Trx) system, which is highly conserved in gram-negative bacteria (3, 10). The Trx system consists of thioredoxin (TrxA) and the NADPH-dependent thioredoxin reductase (TrxB).Any imbalance in the cellular redox state caused by ROS elicits expression of a repertoire of different proteins, commonly under the control of a redox-sensing regulator: for example, OxyR in Escherichia coli and PerR, OhrR, SarZ, and Spx in B. subtilis and S. aureus, respectively (11, 12, 41, 55, 58, 64-66). The subsequently induced proteins detoxify ROS and restore and protect the normal physiological redox state in the cell.Besides ROS and reactive nitrogen species, so-called “reactive electrophilic species” (RES) affect the thiol redox balance. RES include different chemical compounds such as aldehydes, quinones, and the azo compound diamide (2, 43, 45, 46, 53, 66). Quinones and aldehydes have electron-deficient centers that result in thiol-(S) alkylation of cysteine. Exposure of cells to diamide induces the oxidative as well as the electrophile stress response in B. subtilis (43, 45, 53). The toxicity of diamide is based on disulfide bond formation (40), which was recently visualized in B. subtilis and S. aureus by the fluorescence alkylation of oxidized thiols (FALKO) assay (32, 64). It was thought that the formation of nonnative inter- and intramolecular disulfide bonds results in damage of proteins.However, more recent findings demonstrate that diamide stress leads also to S thiolations: formation of disulfide bonds between proteins and LMW thiols (8, 13, 33). S thiolations prevent protein thiols from irreversible oxidation to sulfinic and sulfonic acids, but also affect enzyme activity (35, 47) and signal transduction (39, 42). In B. subtilis, we have identified a few cytoplasmic proteins that are S cysteinylated (33). In addition, the organic peroxide sensor OhrR was inactivated by an S bacillithiolation in B. subtilis (42).Cysteine, BSH, and CoA are also dominant LMW thiols in S. aureus (52). In this study, we have investigated in more detail the extents of S thiolations and inter- and intramolecular disulfide bond formation of B. subtilis and S. aureus in response to disulfide stress. The results showed that exposure to diamide leads to S thiolations in S. aureus. Using a nonreducing/reducing sodium dodecyl sulfate (SDS) diagonal electrophoresis approach, proteins with intermolecular disulfide bonds could be distinguished from proteins with intramolecular disulfide bonds (57). The results support that the majority of reversible thiol oxidations are based on S thiolations rather than disulfide bonds between proteins. Depletion of the free cysteine pool in B. subtilis after exposure to diamide supports this finding. To assess if GSH may have a bearing on the thiol redox buffer of B. subtilis, the gshF gene of Listeria monocytogenes (gshFLm) was expressed in B. subtilis, enabling GSH biosynthesis (29). Although GSH production does not enhance the resistance to oxidative stress in B. subtilis, it participates in the formation of S thiolations.  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