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 441 毫秒
1.
A polyolefinic hydrocarbon was found in nonpolar extracts of Shewanella oneidensis MR-1 and identified as 3,6,9,12,15,19,22,25,28-hentriacontanonaene (compound I) by mass spectrometry, chemical modification, and nuclear magnetic resonance spectroscopy. Compound I was shown to be the product of a head-to-head fatty acid condensation biosynthetic pathway dependent on genes denoted as ole (for olefin biosynthesis). Four ole genes were present in S. oneidensis MR-1. Deletion of the entire oleABCD gene cluster led to the complete absence of nonpolar extractable products. Deletion of the oleC gene alone generated a strain that lacked compound I but produced a structurally analogous ketone. Complementation of the oleC gene eliminated formation of the ketone and restored the biosynthesis of compound I. A recombinant S. oneidensis strain containing oleA from Stenotrophomonas maltophilia strain R551-3 produced at least 17 related long-chain compounds in addition to compound I, 13 of which were identified as ketones. A potential role for OleA in head-to-head condensation was proposed. It was further proposed that long-chain polyunsaturated compounds aid in adapting cells to a rapid drop in temperature, based on three observations. In S. oneidensis wild-type cells, the cellular concentration of polyunsaturated compounds increased significantly with decreasing growth temperature. Second, the oleABCD deletion strain showed a significantly longer lag phase than the wild-type strain when shifted to a lower temperature. Lastly, compound I has been identified in a significant number of bacteria isolated from cold environments.Currently, there is industrial interest in nongaseous microbial hydrocarbons for specialty chemical applications and, more recently, as high-energy biofuels (20, 27, 34). Microbes produce hydrocarbons of different types, for example, aliphatic isoprenoid compounds (20) and alkanes from fatty aldehyde decarbonylation (10). Fatty aldehyde decarbonylation is not well understood but offers a clean route to diesel fuels from fatty acids.Certain microbes also make a distinctly different class of long-chain hydrocarbons, generally C25 to C33 in chain length, that contain a double bond near the middle of the chain (1, 3, 5, 15, 30, 31, 33, 34). These long-chain olefinic hydrocarbons are thought to derive from processes different than isoprene condensation and decarbonylation mechanisms. This class of hydrocarbons has been shown by carbon-14-labeling studies (2) to derive from fatty acids. The process, described in 1929 by Channon and Chibnall (9), has become known as head-to-head hydrocarbon biosynthesis. Albro and Ditmar (3) defined the head-to-head condensation as coupling of the head (C1) and the α-carbon (C2) of two fatty acids with decarboxylation, a reaction that should not be confused with an acyloin-like carboxyl carbon-to-carboxyl carbon coupling. Products of the head-to-head mechanism have been identified in Gram-positive bacteria such as Micrococcus luteus (29, 30) and Arthrobacter aurescens (13) and in Gram-negative bacteria such as Stenotrophomonas maltophilia (28). Micrococcus and Arthrobacter strains produce fatty acids that are methyl branched terminally and subterminally (8, 29, 30). The long-chain olefinic hydrocarbons from those strains similarly contain a mixture of terminal and subterminal methyl group branching (2, 13, 31).Albro and Ditmar (3, 4) acquired direct evidence for the head-to-head mechanism occurring in microbial whole organisms and cell extracts. In cell extracts, it was shown that one of the fatty acid carboxyl groups is lost as carbon dioxide, with the remaining carbon atoms being retained in the resultant hydrocarbon (4). The hydrocarbons contain a double bond at the point of condensation. More recently, Beller et al. described the genes encoding head-to-head fatty acid condensation pathway enzymes from Micrococcus luteus, which are known as ole genes for the olefin products formed (5). Three genes from Micrococcus luteus were shown to confer on Escherichia coli the ability to make long-chain olefinic hydrocarbons. Two recent patent applications by L. Friedman et al. (18 September 2008, WO2008/113041; 4 December 2008, WO2008/147781) also described a three- or four-gene cluster as being involved in head-to-head hydrocarbon biosynthesis to make olefins. The patent applications identified homologs to ole genes in different bacteria, including strains of Shewanella.Bacteria of the genus Shewanella have been heavily studied over the last decade because they are widespread and have the ability to use a startling variety of electron acceptors for respiration (11). There are more than 20 completed genome sequences for Shewanella strains. The model system for studying Shewanella is S. oneidensis MR-1. The genome sequencing of S. oneidensis MR-1 was reported in 2002 (16), and the organism has been shown to be highly amenable to genetic manipulation (11).The present study used Shewanella oneidensis strain MR-1 as a model system to investigate hydrocarbon biosynthetic genes and the possible biological function of the proteins they encode. The hydrocarbon produced by the Ole proteins in S. oneidensis MR-1 was found to be very different from hydrocarbons previously identified as deriving from a head-to-head condensation mechanism (28, 29, 32). The product was identified here as 3,6,9,12,15,19,22,25,28-hentriacontanonaene by chemical modification studies, mass spectrometry, and nuclear magnetic resonance (NMR) spectroscopy. Previously, a similar polyolefin had been identified in many Antarctic bacteria (22). Cloning of a heterologous oleA gene into S. oneidensis MR-1 was found to produce a completely different set of products. A hydrocarbon deletion mutant showed a distinctly longer growth lag than wild-type cells when shifted to a lower temperature, suggesting that the ole genes in S. oneidensis MR-1 may aid the cells in adapting to a sudden drop in temperature.  相似文献   

2.
The single polar flagellum of Shewanella oneidensis MR-1 is powered by two different stator complexes, the sodium-dependent PomAB and the proton-driven MotAB. In addition, Shewanella harbors two genes with homology to motX and motY of Vibrio species. In Vibrio, the products of these genes are crucial for sodium-dependent flagellar rotation. Resequencing of S. oneidensis MR-1 motY revealed that the gene does not harbor an authentic frameshift as was originally reported. Mutational analysis demonstrated that both MotX and MotY are critical for flagellar rotation of S. oneidensis MR-1 for both sodium- and proton-dependent stator systems but do not affect assembly of the flagellar filament. Fluorescence tagging of MotX and MotY to mCherry revealed that both proteins localize to the flagellated cell pole depending on the presence of the basal flagellar structure. Functional localization of MotX requires MotY, whereas MotY localizes independently of MotX. In contrast to the case in Vibrio, neither protein is crucial for the recruitment of the PomAB or MotAB stator complexes to the flagellated cell pole, nor do they play a major role in the stator selection process. Thus, MotX and MotY are not exclusive features of sodium-dependent flagellar systems. Furthermore, MotX and MotY in Shewanella, and possibly also in other genera, must have functions beyond the recruitment of the stator complexes.Flagellum-mediated swimming motility is a widespread means of locomotion among bacteria. Flagella consist of protein filaments that are rotated at the filament''s base by a membrane-embedded motor (3, 39). Rotation is powered by electrochemical gradients across the cytoplasmic membrane. Thus far, two coupling ions, sodium ions and protons, have been described as energy sources for bacterial flagellar motors (4, 24, 48). Two major components confer the conversion of the ion flux into rotary motion. The first component forms a rotor-mounted ring-like structure at the base of the flagellar basal body and is referred to as the switch complex or the C ring; it is composed of the proteins FliG, FliM, and FliN. The second major component is the stator system, consisting of membrane-embedded stator complexes that surround the C ring (3). Each stator complex is composed of two subunits in a 4:2 stoichiometry. In Escherichia coli, MotA and MotB constitute the stator complex by forming a proton-specific ion channel; the Na+-dependent counterpart in Vibrio species consists of the orthologs PomA and PomB (1, 5, 49). MotA and PomA both have four transmembrane domains and are thought to interact with FliG via a cytoplasmic segment to generate torque (2, 50). Stator function is presumably made possible by a peptidoglycan-binding motif located at the C-terminal portion of MotB and PomB that anchors the stator complex to the cell wall (1, 8). In E. coli, at least 11 stator complexes can be synchronously involved in driving flagellar rotation (35). However, a single complex is sufficient for rotation of the filament (36, 40). Despite its tight attachment to the peptidoglycan, the stator ring system was found to form a surprisingly dynamic complex. It has been suggested that inactive precomplexes of the stators form a membrane-located pool before being activated upon incorporation into the stator ring system around the motor (13, 45). In E. coli, the turnover time of stator complexes can be as short as 30 s (21).In Vibrio species, two auxiliary proteins, designated MotX and MotY, are required for motor function of the Na+-driven polar flagellar system (22, 23, 28, 31). Recently, it was shown that the proteins associate with the flagellar basal body in Vibrio alginolyticus to form an additional structure, the T ring (42). MotX interacts with MotY and the PomAB stator complexes, and both proteins are thought to be crucial for the acquisition of the stators to the motor of the polar flagellum. (29, 30, 42). A MotY homolog is also associated with the proton-dependent motor system of the lateral flagella of V. alginolyticus that is induced under conditions of elevated viscosity (41).We recently showed that Shewanella oneidensis MR-1 uses two different stator systems to drive the rotation of its single polar flagellum, the Na+-dependent PomAB stator and the proton-driven MotAB stator. As suggested by genetic data, the MotAB stator has been acquired by lateral gene transfer, presumably in the process of adaptation from a marine to a freshwater environment (32). The two different stators are recruited to the motor in a way that depends on the sodium ion concentration in the medium. The Na+-dependent PomAB stator is present at the flagellated cell pole regardless of the sodium ion concentration, whereas the proton-dependent MotAB stator functionally localizes only under conditions of low sodium or in the absence of PomAB. It is still unclear how stator selection is achieved and whether additional proteins play a role in this process.Orthologs of motX and motY have been annotated in S. oneidensis MR-1. We thus hypothesized that MotX and MotY might play a role in stator selection in S. oneidensis MR-1. However, the originally published sequence of motY harbors a frameshift that would result in a drastically truncated protein lacking a functionally relevant putative peptidoglycan-binding domain at its C terminus (16, 18). This situation seemed inconsistent with a role for MotY in S. oneidensis MR-1.Here we describe a functional analysis of the MotX and MotY orthologs in S. oneidensis MR-1. We found that motY does not, in fact, contain a frameshift mutation, so that MotY is translated in its full-length form. Both MotX and MotY were essential for Na+-dependent and proton-dependent motility. Therefore, these proteins have a role in S. oneidensis MR-1 that differs from their function in Vibrio species. We also used fusions to the fluorescent protein mCherry for functional localization studies of MotX and MotY.  相似文献   

3.
Although members of the genus Shewanella have common features (e.g., the presence of decaheme c-type cytochromes [c-cyts]), they are widely variable in genetic and physiological features. The present study compared the current-generating ability of S. loihica PV-4 in microbial fuel cells (MFCs) with that of well-characterized S. oneidensis MR-1 and examined the roles of c-cyts in extracellular electron transfer. We found that strains PV-4 and MR-1 exhibited notable differences in current-generating mechanisms. While the MR-1 MFCs maintained a constant current density over time, the PV-4 MFCs continued to increase in current density and finally surpassed the MR-1 MFCs. Coulombic efficiencies reached 26% in the PV-4 MFC but 16% in the MR-1 MFCs. Although both organisms produced quinone-like compounds, anode exchange experiments showed that anode-attached cells of PV-4 produced sevenfold more current than planktonic cells in the same chamber, while planktonic cells of MR-1 produced twice the current of the anode-attached cells. Examination of the genome sequence indicated that PV-4 has more c-cyt genes in the metal reductase-containing locus than MR-1. Mutational analysis revealed that PV-4 relied predominantly on a homologue of the decaheme c-cyt MtrC in MR-1 for current generation, even though it also possesses two homologues of the decaheme c-cyt OmcA in MR-1. These results suggest that current generation in a PV-4 MFC is in large part accomplished by anode-attached cells, in which the MtrC homologue constitutes the main path of electrons toward the anode.Some species of dissimilatory metal-reducing bacteria (DMRB) are able to reduce solid metal oxides as terminal electron acceptors and generate currents in microbial fuel cells (MFCs) (2, 11, 14, 30, 46). Although mixed cultures are often used in MFC experiments (13), studies seeking a mechanistic understanding of electron transfer to electrode surfaces typically target pure cultures of such DMRB, due to the complexity in microbial communities. Presently, two model DMRB, Shewanella oneidensis MR-1 and Geobacter sulfurreducens PCA (2, 3, 12, 18, 31), are used in most investigations.S. oneidensis MR-1 is a metabolically diverse DMRB that has been studied extensively for its potential use in bioremediation applications. For this reason, MR-1 was the first Shewanella species to have its genome completely sequenced and annotated (10). In addition, since the first report in 1999 when this microorganism was shown to have the ability to transfer electrons to the electrode without an exogenously added mediator (14), it has also become one of the model organisms for the study of electron transfer mechanisms in MFCs.Although the molecular mechanisms for extracellular electron transfer have not yet been elucidated fully, c-type cytochromes (c-cyts) appear to be the key cellular components involved in this process (38). In S. oneidensis MR-1, OmcA and MtrC are outer membrane (OM), decaheme c-cyts that are considered to be involved in the direct (directly attached) electron transfer to solid metal oxides and anodes of MFCs (9, 20, 22, 23, 47). Several pieces of evidence suggest that OmcA and MtrC form a complex and act in a cooperative manner (33, 37, 42), and these results correlate with the fact that the genes encoding these proteins constitute an operon-like cluster in the chromosome (1). It has also been shown that MtrC and OmcA have overlapping functions as terminal reductases of metal oxides (25, 38). OmcA and MtrC are also present on the surface of nanowires and may be involved in the long-range transfer of electrons (8). In addition to direct electron transfer, MR-1 has the ability to produce water-soluble electron-shuttle compounds (quinones and flavins) that are involved in the mediated electron transfer from cells to distant solid electron acceptors (metal oxides or MFC anodes) (21, 27, 44).Recently, the genome sequences of nearly 20 Shewanella strains have been completed and annotated, opening the door to study the diversity of their extracellular electron transfer mechanisms. A comparison of their genomes has shown that although they have some consensus OM c-cyt genes, variations exist in the number and order of these genes in their metal reductase-containing loci (6). One such species is S. loihica strain PV-4, which was recently isolated from an iron-rich microbial mat near a deep-sea hydrothermal vent located on the Loihi Seamount in Hawaii (7, 32). The phenotypic and phylogenetic characteristics of PV-4 were determined, with a subsequent study focusing on the metal reduction and iron biomineralization capabilities of this bacterium (32). Initial experiments performed in our laboratory revealed that PV-4 developed a c-cyt-dependent deep red color that was much more striking than that of strain MR-1 when grown anaerobically with iron oxide as the terminal electron acceptor (26). This allowed us to assume that PV-4 could have a high extracellular electron transfer ability. Accordingly, the present study evaluated the current-producing ability of strain PV-4 in MFCs and examined the roles of some c-cyts in extracellular electron transfer. Special attention was paid to the comparison of PV-4 with MR-1 to reveal differences in mechanisms for extracellular electron transfer. We report herein differences between these strains in the roles of OM c-cyts for extracellular electron transfer, the behaviors and metabolic patterns of MFC, and the resultant MFC performances.  相似文献   

4.
The Mtr respiratory pathway of Shewanella oneidensis strain MR-1 is required to effectively respire both soluble and insoluble forms of oxidized iron. Flavins (riboflavin and flavin mononucleotide) recently have been shown to be excreted by MR-1 and facilitate the reduction of insoluble substrates. Other Shewanella species tested accumulated flavins in supernatants to an extent similar to that of MR-1, suggesting that flavin secretion is a general trait of the species. External flavins have been proposed to act as both a soluble electron shuttle and a metal chelator; however, at biologically relevant concentrations, our results suggest that external flavins primarily act as electron shuttles for MR-1. Using deletion mutants lacking various Mtr-associated proteins, we demonstrate that the Mtr extracellular respiratory pathway is essential for the reduction of flavins and that decaheme cytochromes found on the outer surface of the cell (MtrC and OmcA) are required for the majority of this activity. Given the involvement of external flavins in the reduction of electrodes, we monitored current production by Mtr respiratory pathway mutants in three-electrode bioreactors under controlled flavin concentrations. While mutants lacking MtrC were able to reduce flavins at 50% of the rate of the wild type in cell suspension assays, these strains were unable to grow into productive electrode-reducing biofilms. The analysis of mutants lacking OmcA suggests a role for this protein in both electron transfer to electrodes and attachment to surfaces. The parallel phenotypes of Mtr mutants in flavin and electrode reduction blur the distinction between direct contact and the redox shuttling strategies of insoluble substrate reduction by MR-1.Shewanella oneidensis strain MR-1 (MR-1) is a facultative anaerobe capable of respiring a variety of substrates, including various metals and metal oxides, a phenotype that is important for bioremediation and metal cycling in natural environments (22, 53). At near-neutral pH, Fe(III) and Mn(IV) often are present as insoluble oxide minerals. Dissimilatory metal-reducing bacteria such as MR-1 have developed pathways to transfer electrons from the interior of the cell to these external terminal electron acceptors. In some bacteria, these pathways also can transfer electrons to electrodes, which can be harnessed for renewable energy and remote biosensor applications (23, 26, 27). Beyond increasing our understanding of this unusual process, applying anaerobic microbial extracellular respiration to new technologies requires a thorough understanding of the molecular dynamics and cellular physiology of electron source utilization (substrate oxidation) and the reduction of insoluble terminal electron acceptor(s). There are four proposed mechanisms to explain how insoluble substrates are reduced by Shewanella: (i) direct contact, (ii) electron shuttling, (iii) chelation, and (iv) electrically conductive appendages (reviewed in reference 18). We will focus on the first three strategies here.Flavins recently have been discovered to accelerate the reduction of both iron oxide minerals (51) and electrodes (30) by MR-1. Riboflavin (vitamin B2) is a precursor for the biosynthesis of flavin mononucleotide (FMN) and flavin adenine dinucleotide (FAD) (13). Riboflavin and FMN both can be observed to build up in the supernatant of anaerobic and aerobically grown cultures of MR-1 (30, 51). However, the mechanism by which flavins enhance the rate of iron oxide mineral or electrode reduction is unknown, although recent work is consistent with a critical role for these compounds in mediating solid Fe(III) reduction by MR-1 (42). Since soluble (chelated) Fe(III) is reduced faster than insoluble Fe(III) by MR-1 (6), one possible explanation for the enhancement of insoluble iron reduction by flavins is increased available soluble iron via chelation (1, 2, 30). Flavins also may be utilized as redox-active compounds to traffic electrons between extracellular reductases on the surface of the cell and insoluble substrates (30, 51), a process termed electron shuttling (18, 39, 41). The chelation of the terminal electron acceptor during electrode reduction is not relevant when the anode is composed of graphite. Therefore, electron shuttling likely is responsible for the flavin enhancement of current production on poised-potential electrodes (30). However, it is unclear if the chelation of metals by flavins influences insoluble metal reduction by S. oneidensis (30).The Mtr pathway is required for the reduction of metals and electrodes (5, 6, 9, 17). Five primary protein components have been identified in this pathway: OmcA, MtrC, MtrA, MtrB, and CymA (47). Current models of electron transfer in MR-1 assume that electrons from carbon source oxidation are passed via the menaquinone pool to the inner membrane-anchored c-type cytochrome CymA (19, 31). These electrons then are transferred to a periplasmic c-type cytochrome, MtrA, and eventually to outer membrane (OM)-anchored c-type cytochromes MtrC and OmcA, which interact with an integral OM scaffolding protein, MtrB (32, 33, 43). These OM cytochromes then can reduce various substrates, including iron oxides and electrodes (8, 9, 12, 36, 47). Since the Mtr system is required by MR-1 to reduce many different substrates (18), it also could be capable of reducing extracellular flavins. Indeed, electron transfer to carbon electrodes is impaired in strains lacking Mtr pathway components (9, 17), which may be explained simply by a decreased ability to reduce extracellular flavins. The observation that Mtr mutants produce less current on electrodes than the wild type could be due to (i) less current generated per cell (either direct reduction or flavin mediated), (ii) decreased attachment to the electrode surface, (iii) differences in external flavin concentrations, or (iv) a combination of these three possibilities. Determining the specific activity (current produced per unit of attached biomass) of Mtr mutants on electrodes under conditions where flavin levels were controlled would allow for differentiation between these possibilities. To date, this kind of analysis has not been reported.The results presented here extend our knowledge of how S. oneidensis catalyzes the reduction of insoluble substrates. Experiments using a model iron chelator and electron shuttle are consistent with electron shuttling being the primary mechanism by which flavins enhance insoluble iron oxide reduction rates. Moreover, we demonstrate that MR-1 reduces extracellular flavins at physiologically relevant rates and that the Mtr pathway accounts for at least 95% of this activity. The specific activities of various mutant strains lacking Mtr pathway components on poised-potential electrodes also are reported. Our data suggest that MtrC is responsible for most of the electron transfer to carbon electrodes, while OmcA is involved in attachment and has a lesser role in electron transfer. These observations could have broader implications regarding the role of OmcA in the reduction of soluble and insoluble substrates (8, 9, 36).  相似文献   

5.
6.
A microbial fuel cell (MFC) was inoculated with a random transposon insertion mutant library of Shewanella oneidensis MR-1 and operated with lactate as the sole fuel to select for mutants that preferentially grew in it. Agar plate cultivation of the resultant MFC enrichment culture detected an increased number of colonies exhibiting rough morphology. One such isolate, strain 4A, generated 50% more current in an MFC than wild-type MR-1. Determination of the transposon insertion site in strain 4A followed by deletion and complementation experiments revealed that the SO3177 gene, encoding a putative formyltransferase and situated in a cell surface polysaccharide biosynthesis gene cluster, was responsible for the increased current. Transmission electron microscopy showed that a layered structure at the cell surface, stainable with ruthenium red, was impaired in the SO3177 mutant (ΔSO3177), confirming that SO3177 is involved in the biosynthesis of cell surface polysaccharides. Compared to the wild type, ΔSO3177 cells preferentially attached to graphite felt anodes in MFCs, while physicochemical analyses revealed that the cell surface of ΔSO3177 was more hydrophobic. These results demonstrate that cell surface polysaccharides affect not only the cell adhesion to graphite anodes but also the current generation in MFCs.Dissimilatory metal-reducing bacteria (DMRB) conserve energy for growth by coupling the oxidation of organic compounds to the reduction of metal compounds (29). DMRB are of great interest not only for their importance in the biogeochemical cycling of metals (25) but also for their utility in biotechnological processes, such as microbial fuel cells (MFCs) (24, 40). In recent years, the ability of many DMRB, including members of the genera Shewanella (5, 12, 20, 31), Geobacter (2), Aeromonas (34), Desulfobulbus (19), and Phodoferax (9), to generate current in MFCs has been described.Among DMRB, Shewanella oneidensis MR-1 is one of the most extensively studied due to its metabolic versatility (28), annotated genome sequence (17), and genetic accessibility. In addition, since the first report in 1999 when this microorganism was shown to have the ability to transfer electrons to an anode without an exogenously added mediator (20), it has become a model organism for the study of microbial current generation in MFCs. Extensive studies have been performed to understand the mechanisms of extracellular electron transfer (EET) to solid materials, such as MFC anodes and metal oxides, in strain MR-1. Multiple mechanisms, including direct EET through the physical contact of bacterial cells via outer membrane (OM) cytochromes (42) and conductive nanowires (16) and mediated EET via self-produced electron shuttles such as quinones and flavins (27, 30, 39, 41), have been identified.Although OM cytochromes and electron shuttles have been identified to play central roles in EET, it is reasonable to speculate that this complex catabolic process is also influenced by other (extra)cellular components. To date, only limited studies have been done to investigate other cellular components involved in EET (7). A useful approach for identifying unknown cellular components (and genes) associated with a particular phenotype involves the construction and screening of a random mutant library for obtaining mutants with altered phenotypes. In the present study, we constructed a random transposon (Tn) insertion mutant library of S. oneidensis MR-1 and obtained mutants with altered colony morphologies (rough morphotypes) after the selection of mutants in an MFC. Analyses of one of such mutants suggest that cell surface capsular polysaccharides affect not only the adhesion of cells to graphite anodes but also the current generation in MFCs.  相似文献   

7.
The ability to undergo dramatic morphological changes in response to extrinsic cues is conserved in fungi. We have used the model yeast Schizosaccharomyces pombe to determine which intracellular signal regulates the dimorphic switch from the single-cell yeast form to the filamentous invasive growth form. The S. pombe Asp1 protein, a member of the conserved Vip1 1/3 inositol polyphosphate kinase family, is a key regulator of the morphological switch via the cAMP protein kinase A (PKA) pathway. Lack of a functional Asp1 kinase domain abolishes invasive growth which is monopolar, while an increase in Asp1-generated inositol pyrophosphates (PP) increases the cellular response. Remarkably, the Asp1 kinase activity encoded by the N-terminal part of the protein is regulated negatively by the C-terminal domain of Asp1, which has homology to acid histidine phosphatases. Thus, the fine tuning of the cellular response to environmental cues is modulated by the same protein. As the Saccharomyces cerevisiae Asp1 ortholog is also required for the dimorphic switch in this yeast, we propose that Vip1 family members have a general role in regulating fungal dimorphism.Eucaryotic cells are able to define and maintain a particular cellular organization and thus cellular morphology by executing programs modulated by internal and external signals. For example, signals generated within a cell are required for the selection of the growth zone after cytokinesis in the fission yeast Schizosaccharomyces pombe or the emergence of the bud in Saccharomyces cerevisiae (37, 44, 81). Cellular morphogenesis is also subject to regulation by a wide variety of external signals, such as growth factors, temperature, hormones, nutrient limitation, and cell-cell or cell-substrate contact (13, 34, 66, 75, 81). Both types of signals will lead to the selection of growth zones accompanied by the reorganization of the cytoskeleton.The ability to alter the growth form in response to environmental conditions is an important virulence-associated trait of pathogenic fungi which helps the pathogen to spread in and survive the host''s defense system (7, 32). Alteration of the growth form in response to extrinsic signals is not limited to pathogenic fungi but is also found in the model yeasts S. cerevisiae and S. pombe, in which it appears to represent a foraging response (1, 24).The regulation of polarized growth and the definition of growth zones have been studied extensively with the fission yeast S. pombe. In this cylindrically shaped organism, cell wall biosynthesis is restricted to one or both cell ends in a cell cycle-regulated manner and to the septum during cytokinesis (38). This mode of growth requires the actin cytoskeleton to direct growth and the microtubule cytoskeleton to define the growth sites (60). In interphase cells, microtubules are organized in antiparallel bundles that are aligned along the long axis of the cell and grow from their plus ends toward the cell tips. Upon contact with the cell end, microtubule growth will first pause and then undergo a catastrophic event and microtubule shrinkage (21). This dynamic behavior of the microtubule plus end is regulated by a disparate, conserved, microtubule plus end group of proteins, called the +TIPs. The +TIP complex containing the EB1 family member Mal3 is required for the delivery of the Tea1-Tea4 complex to the cell tip (6, 11, 27, 45, 77). The latter complex docks at the cell end and recruits proteins required for actin nucleation (46, 76). Thus, the intricate cross talk between the actin and the microtubule cytoskeleton at specific intracellular locations is necessary for cell cycle-dependent polarized growth of the fission yeast cell.The intense analysis of polarized growth control in single-celled S. pombe makes this yeast an attractive organism for the identification of key regulatory components of the dimorphic switch. S. pombe multicellular invasive growth has been observed for specific strains under specific conditions, such as nitrogen and ammonium limitation and the presence of excess iron (1, 19, 50, 61).Here, we have identified an evolutionarily conserved key regulator of the S. pombe dimorphic switch, the Asp1 protein. Asp1 belongs to the highly conserved family of Vip1 1/3 inositol polyphosphate kinases, which is one of two families that can generate inositol pyrophosphates (PP) (17, 23, 42, 54). The inositol polyphosphate kinase IP6K family, of which the S. cerevisiae Kcs1 protein is a member, is the “classical” family that can phosphorylate inositol hexakisphosphate (IP6) (70, 71). These enzymes generate a specific PP-IP5 (IP7), which has the pyrophosphate at position 5 of the inositol ring (20, 54). The Vip1 family kinase activity was unmasked in an S. cerevisiae strain with KCS1 and DDP1 deleted (54, 83). The latter gene encodes a nudix hydrolase (14, 68). The mammalian and S. cerevisiae Vip1 proteins phosphorylate the 1/3 position of the inositol ring, generating 1/3 diphosphoinositol pentakisphosphate (42). Both enzyme families collaborate to generate IP8 (17, 23, 42, 54, 57).Two modes of action have been described for the high-energy moiety containing inositol pyrophosphates. First, these molecules can phosphorylate proteins by a nonenzymatic transfer of a phosphate group to specific prephosphorylated serine residues (2, 8, 69). Second, inositol pyrophosphates can regulate protein function by reversible binding to the S. cerevisiae Pho80-Pho85-Pho81 complex (39, 40). This cyclin-cyclin-dependent kinase complex is inactivated by inositol pyrophosphates generated by Vip1 when cells are starved of inorganic phosphate (39, 41, 42).Regulation of phosphate metabolism in S. cerevisiae is one of the few roles specifically attributed to a Vip1 kinase. Further information about the cellular function of this family came from the identification of the S. pombe Vip1 family member Asp1 as a regulator of the actin nucleator Arp2/3 complex (22). The 106-kDa Asp1 cytoplasmic protein, which probably exists as a dimer in vivo, acts as a multicopy suppressor of arp3-c1 mutants (22). Loss of Asp1 results in abnormal cell morphology, defects in polarized growth, and aberrant cortical actin cytoskeleton organization (22).The Vip1 family proteins have a dual domain structure which consists of an N-terminal “rimK”/ATP-grasp superfamily domain found in certain inositol signaling kinases and a C-terminal part with homology to histidine acid phosphatases present in phytase enzymes (28, 53, 54). The N-terminal domain is required and sufficient for Vip1 family kinase activity, and an Asp1 variant with a mutation in a catalytic residue of the kinase domain is unable to suppress mutants of the Arp2/3 complex (17, 23, 54). To date, no function has been described for the C-terminal phosphatase domain, and this domain appears to be catalytically inactive (17, 23, 54).Here we describe a new and conserved role for Vip1 kinases in regulating the dimorphic switch in yeasts. Asp1 kinase activity is essential for cell-cell and cell-substrate adhesion and the ability of S. pombe cells to grow invasively. Interestingly, Asp1 kinase activity is counteracted by the putative phosphatase domain of this protein, a finding that allows us to describe for the first time a function for the C-terminal part of Vip1 proteins.  相似文献   

8.
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   

9.
10.
11.
12.
13.
14.
15.
16.
Claudin-1, a component of tight junctions between liver hepatocytes, is a hepatitis C virus (HCV) late-stage entry cofactor. To investigate the structural and functional roles of various claudin-1 domains in HCV entry, we applied a mutagenesis strategy. Putative functional intracellular claudin-1 domains were not important. However, we identified seven novel residues in the first extracellular loop that are critical for entry of HCV isolates drawn from six different subtypes. Most of the critical residues belong to the highly conserved claudin motif W30-GLW51-C54-C64. Alanine substitutions of these residues did not impair claudin-1 cell surface expression or lateral protein interactions within the plasma membrane, including claudin-1-claudin-1 and claudin-1-CD81 interactions. However, these mutants no longer localized to cell-cell contacts. Based on our observations, we propose that cell-cell contacts formed by claudin-1 may generate specialized membrane domains that are amenable to HCV entry.Hepatitis C virus (HCV) is a major human pathogen that affects approximately 3% of the global population, leading to cirrhosis and hepatocellular carcinoma in chronically infected individuals (5, 23, 42). Hepatocytes are the major target cells of HCV (11), and entry follows a complex cascade of interactions with several cellular factors (6, 8, 12, 17). Infectious viral particles are associated with lipoproteins and initially attach to target cells via glycosaminoglycans and the low-density lipoprotein receptor (1, 7, 31). These interactions are followed by direct binding of the E2 envelope glycoprotein to the scavenger receptor class B type I (SR-B1) and then to the CD81 tetraspanin (14, 15, 33, 36). Early studies showed that CD81 and SR-B1 were necessary but not sufficient for HCV entry, and claudin-1 was discovered to be a requisite HCV entry cofactor that appears to act at a very late stage of the process (18).Claudin-1 is a member of the claudin protein family that participates in the formation of tight junctions between adjacent cells (25, 30, 37). Tight junctions regulate the paracellular transport of solutes, water, and ions and also generate apical-basal cell polarity (25, 37). In the liver, the apical surfaces of hepatocytes form bile canaliculi, whereas the basolateral surfaces face the underside of the endothelial layer that lines liver sinusoids. Claudin-1 is highly expressed in tight junctions formed by liver hepatocytes as well as on all hepatoma cell lines that are permissive to HCV entry (18, 24, 28). Importantly, nonhepatic cell lines that are engineered to express claudin-1 become permissive to HCV entry (18). Claudin-6 and -9 are two other members of the human claudin family that enable HCV entry into nonpermissive cells (28, 43).The precise role of claudin-1 in HCV entry remains to be determined. A direct interaction between claudins and HCV particles or soluble E2 envelope glycoprotein has not been demonstrated (18; T. Dragic, unpublished data). It is possible that claudin-1 interacts with HCV entry receptors SR-B1 or CD81, thereby modulating their ability to bind to E2. Alternatively, claudin-1 may ferry the receptor-virus complex to fusion-permissive intracellular compartments. Recent studies show that claudin-1 colocalizes with the CD81 tetraspanin at the cell surface of permissive cell lines (22, 34, 41). With respect to nonpermissive cells, one group observed that claudin-1 was predominantly intracellular (41), whereas another reported associations of claudin-1 and CD81 at the cell surface, similar to what is observed in permissive cells (22).Claudins comprise four transmembrane domains along with two extracellular loops and two cytoplasmic domains (19, 20, 25, 30, 37). The first extracellular loop (ECL1) participates in pore formation and influences paracellular charge selectivity (25, 37). It has been shown that the ECL1 of claudin-1 is required for HCV entry (18). All human claudins comprise a highly conserved motif, W30-GLW51-C54-C64, in the crown of ECL1 (25, 37). The exact function of this domain is unknown, and we hypothesized that it is important for HCV entry. The second extracellular loop is required for the holding function and oligomerization of the protein (25). Claudin-1 also comprises various signaling domains and a PDZ binding motif in the intracellular C terminus that binds ZO-1, another major component of tight junctions (30, 32, 37). We further hypothesized that some of these domains may play a role in HCV entry.To understand the role of claudin-1 in HCV infection, we developed a mutagenesis strategy targeting the putative sites for internalization, glycosylation, palmitoylation, and phosphorylation. The functionality of these domains has been described by others (4, 16, 25, 35, 37, 40). We also mutagenized charged and bulky residues in ECL1, including all six residues within the highly conserved motif W30-GLW51-C54-C64. None of the intracellular domains were found to affect HCV entry. However, we identified seven residues in ECL1 that are critical for entry mediated by envelope glycoproteins derived from several HCV subtypes, including all six residues of the conserved motif. These mutants were still expressed at the cell surface and able to form lateral homophilic interactions within the plasma membrane as well as to engage in lateral interactions with CD81. In contrast, they no longer engaged in homophilic trans interactions at cell-cell contacts. We conclude that the highly conserved motif W30-GLW51-C54-C64 of claudin-1 is important for HCV entry into target cells and participates in the formation of cell-cell contacts.  相似文献   

17.
18.
Shewanella oneidensis is able to conserve energy for growth by reducing a wide variety of terminal electron acceptors during anaerobic respiration, including several environmentally hazardous pollutants. This bacterium employs various electron transfer mechanisms for anaerobic respiration, including cell-bound reductases and secreted redox mediators. The aim of this study was to develop rapid tools for profiling the key metabolic changes associated with these different growth regimes and physiological responses. Initial experiments focused on comparing cells grown under aerobic and anaerobic conditions. Fourier transform infrared (FT-IR) spectroscopy with cluster analysis showed that there were significant changes in the metabolic fingerprints of the cells grown under these two culture conditions. FT-IR spectroscopy clearly differentiated cells of S. oneidensis MR-1 cultured at various growth points and cells grown using different electron acceptors, resulting in different phenotypic trajectories in the cluster analysis. This growth-related trajectory analysis is applied successfully for the first time, here with FT-IR spectroscopy, to investigate the phenotypic changes in contrasting S. oneidensis cells. High-performance liquid chromatography (HPLC) was also used to quantify the concentrations of flavin compounds, which have been identified recently as extracellular redox mediators released by a range of Shewanella species. The partial least-squares regression (PLSR) multivariate statistical technique was combined with FT-IR spectroscopy to predict the concentrations of the flavins secreted by cells of S. oneidensis MR-1, suggesting that this combination could be used as a rapid alternative to conventional chromatographic methods for analysis of flavins in cell cultures. Furthermore, coupling of the FT-IR spectroscopy and HPLC techniques appears to offer a potentially useful tool for rapid characterization of the Shewanella cell metabolome in various process environments.Shewanella oneidensis, a Gram-negative dissimilatory metal-reducing bacterium, is able to conserve energy for growth by reducing a variety of terminal electron acceptors during aerobic and anaerobic respiration (24), including several environmentally hazardous pollutants (1). The terminal electron acceptors used during anaerobic respiration vary, ranging from nitrate, fumarate, trimethylamine N-oxide (TMAO), dimethyl sulfoxide (DMSO), and sulfur compounds to fuel cell anodes and various metals and metalloids, including insoluble metal oxides (5, 33, 34).There has been intense interest in the versatile metabolism of this bacterium and its potential to respire and bioremediate toxic environmental chemicals and metals, such as U(VI) and Cr(VI) (25, 43), under anoxic conditions. Riboflavin (RF), flavin mononucleotide (FMN), and flavin-adenine dinucleotide (FAD) were identified as the dominant electron shuttles secreted by a diversity of Shewanella cells and shown to mediate extracellular reduction of insoluble Fe(III) minerals and organic molecules, including azo dyes (23, 40). Field and Brady tested riboflavin and found that its presence during anaerobic reduction of azo dyes improves the overall kinetics of the reduction process (9). von Canstein et al. recently found that Shewanella cells are able to secrete flavins, FAD, FMN, and riboflavin as extracellular redox mediators and quantified these using high-performance liquid chromatography-mass spectrometry (HPLC-MS) (44). The experiments showed that the production profiles of these three chemicals were different under anaerobic and aerobic regimes, and this can be used to distinguish the two metabolic pathways with oxygen or fumarate as electron acceptors.Although we are gaining a deeper understanding of the genetic and biochemical basis of the diverse respiration pathways of this organism, supported by the recent availability of the complete genome sequence (14), complementary metabolomic approaches have not been used to identify or quantify the metabolic changes expected with major physiological shifts in this organism under contrasting growth regimes. Fourier transform infrared (FT-IR) spectroscopy was chosen for this study, as this method offers the advantage of minimal sample preparation, as well as being rapid, nondestructive, readily automatable, relatively inexpensive, and quantitative, compared with other metabolic profiling techniques (8, 12, 15, 35). The infrared absorbance spectra generated by FT-IR spectroscopy have been used to identify specific biochemical features and also provide a global biochemical “fingerprint” for mixed, complex samples (11, 27) in many research areas (10, 11). For instance, FT-IR spectroscopy has proved sensitive enough for analysis of the chemical composition of a single strain of Escherichia coli after exposure to ionic liquids (6), for functional genomics screening (22), for measuring abiotic perturbations in algae (39), and for characterization of microbial degradation pathways (16). We believe that this technique offers considerable potential for rapid differentiation of metabolic changes of bacteria responding to contrasting growth regimes in natural and engineered environments, although this remains to be demonstrated.The FT-IR spectra generated from microorganisms have very complex profiles that can be related back to biochemical components which one would expect to be detected in the samples. While some subtle quantitative differences can be observed between the spectra, there are few, if any, qualitative differences to be seen, and so it is almost impossible to interpret these data with the naked eye (19, 20, 45). Therefore, multivariate statistical techniques are needed to model the relationship between the phenotypic changes occurring in this organism during growth with the observations recorded by FT-IR spectroscopy.In the present study, we investigated the ability of FT-IR spectroscopy to analyze the S. oneidensis MR-1 cell metabolome and distinguish significant metabolic changes associated with anaerobic and aerobic growth conditions. Analyses were conducted directly on the cells for the endometabolome in assessing the metabolic fingerprint. Both unsupervised and supervised learning methods (viz., principal component analysis [PCA] [21] and discriminant function analysis [DFA] [26]) were used to identify any differences between the FT-IR spectra from cells grown under the two regimes, and the contributions of a range of biomolecules were elucidated. In addition to metabolic fingerprinting of whole cells by FT-IR, key components of the extracellular metabolome (the so-called exometabolome, or metabolic footprint) were also quantified by HPLC analysis. Here, riboflavin, FMN, and FAD in culture supernatants were quantified, and partial least-squares regression (PLSR) was used to identify correlations between the cell''s FT-IR spectra and flavin concentrations assessed using HPLC data. These studies show for the first time that FT-IR analyses can be used for rapid identification of metabolic shifts in both the intracellular and the extracellular metabolomes of S. oneidensis MR-1 cultures.  相似文献   

19.
Human immunodeficiency virus type 1 (HIV-1) envelope protein (Env) is subject to both neutralizing antibody (NAb) and CD8 T-cell (cytotoxic T-lymphocyte [CTL]) immune pressure. We studied the reversion of the Env CTL escape mutant virus to the wild type and the relationship between the reversion of CTL mutations with N-linked glycosylation site (NLGS)-driven NAb escape in pigtailed macaques. Env CTL mutations either did not revert to the wild type or only transiently reverted 5 to 7 weeks after infection. The CTL escape mutant reversion was coincident, for the same viral clones, with the loss of NLGS mutations. At one site studied, both CTL and NLGS mutations were needed to confer NAb escape. We conclude that CTL and NAb escape within Env can be tightly linked, suggesting opportunities to induce effective multicomponent anti-Env immunity.CD8 T-cell responses against human immunodeficiency virus (HIV) have long been observed to select for viral variants that avoid cytotoxic T-lymphocyte (CTL) recognition (2, 5, 15, 18, 27). These immune escape mutations may, however, result in reduced replication competence (“fitness cost”) (11, 20, 26). CTL escape variants have been shown to revert to the wild type (WT) upon passage to major histocompatibility complex-mismatched hosts, both in macaques with simian immunodeficiency virus (SIV) or chimeric SIV/HIV (SHIV) infection (11, 12) and in humans with HIV type 1 (HIV-1) infection (1, 19).Most analyses of CTL escape and reversion have studied Gag CTL epitopes known to facilitate control of viremia (7, 14, 21, 30). Fewer analyses have studied Env-specific CTL epitopes. Recent sequencing studies suggest the potential for mutations within predicted HIV-1 Env-specific CTL epitopes to undergo reversion to the WT (16, 23). Env-specific CTL responses may, however, have less impact on viral control of both HIV-1 and SIV/SHIV than do Gag CTL responses (17, 24, 25), presumably reflecting either less-potent inhibition of viral replication or minimal fitness cost of escape (9).Serial viral escape from antibody pressure also occurs in both macaques and humans (3, 13, 28). Env is extensively glycosylated, and this “evolving glycan shield” can sterically block antibody binding without mutation at the antibody-binding site (8, 16, 31). Mutations at glycosylation sites, as well as other mutations, are associated with escape from neutralizing antibody (NAb) responses (4, 13, 29). Mutations in the amino acid sequences of N-linked glycosylation sites (NLGS) can alter the packing of the glycan cloud that surrounds the virion, by a loss, gain, or shift of an NLGS (32), thus facilitating NAb escape.Env is the only viral protein targeted by both CTL and NAb responses. The serial viral escape from both Env-specific CTL and NAb responses could have implications for viral fitness and the reversion of multiple mutations upon transmission to naïve hosts.We previously identified three common HIV-1 Env-specific CD8 T cell epitopes, RY8788-795, SP9110-118, and NL9671-679, and their immune escape patterns in pigtail macaques (Macaca nemestrina) infected with SHIVmn229 (25). SHIVmn229 is a chimeric virus constructed from an SIVmac239 backbone and an HIV-1HXB2 env fragment that was passaged through macaques to become pathogenic (11). This earlier work provided an opportunity for detailed studies of how viruses with Env-specific CTL escape mutations, as well as mutations in adjacent NLGS, evolve when transmitted to naïve pigtail macaques.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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