首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The biological, serological, and genomic characterization of a paramyxovirus recently isolated from rockhopper penguins (Eudyptes chrysocome) suggested that this virus represented a new avian paramyxovirus (APMV) group, APMV10. This penguin virus resembled other APMVs by electron microscopy; however, its viral hemagglutination (HA) activity was not inhibited by antisera against any of the nine defined APMV serotypes. In addition, antiserum generated against this penguin virus did not inhibit the HA of representative viruses of the other APMV serotypes. Sequence data produced using random priming methods revealed a genomic structure typical of APMV. Phylogenetic evaluation of coding regions revealed that amino acid sequences of all six proteins were most closely related to APMV2 and APMV8. The calculation of evolutionary distances among proteins and distances at the nucleotide level confirmed that APMV2, APMV8, and the penguin virus all were sufficiently divergent from each other to be considered different serotypes. We propose that this isolate, named APMV10/penguin/Falkland Islands/324/2007, be the prototype virus for APMV10. Because of the known problems associated with serology, such as antiserum cross-reactivity and one-way immunogenicity, in addition to the reliance on the immune response to a single protein, the hemagglutinin-neuraminidase, as the sole base for viral classification, we suggest the need for new classification guidelines that incorporate genome sequence comparisons.Viruses from the Paramyxoviridae family have caused disease in humans and animals for centuries. Over the last 40 years, many paramyxoviruses isolated from animals and people have been newly described (16, 17, 22, 29, 31, 32, 36, 42, 44, 46, 49, 58, 59, 62-64). Viruses from this family are pleomorphic, enveloped, single-stranded, nonsegmented, negative-sense RNA viruses that demonstrate serological cross-reactivity with other paramyxoviruses related to them (30, 46). The subfamily Paramyxovirinae is divided into five genera: Respirovirus, Morbillivirus, Rubulavirus, Henipavirus, and Avulavirus (30). The Avulavirus genus contains nine distinct avian paramyxovirus (APMV) serotypes (Table (Table1),1), and information on the discovery of each has been reported elsewhere (4, 6, 7, 9, 12, 34, 41, 50, 51, 60, 68).

TABLE 1.

Characteristics of prototype viruses APMV1 to APMV9 and the penguin virus
StrainHostDiseaseDistributionFusion cleavagecGI accession no.
APMV1/Newcastle disease virus>250 speciesHigh mortalityWorldwideGRRQKRF45511218
InapparentWorldwideGGRQGRLa11545722
APMV2/Chicken/CA/Yucaipa/1956Turkey, chickens, psittacines, rails, passerinesDecrease in egg production and respiratory diseaseWorldwideDKPASRF169144527
APMV3/Turkey/WI/1968TurkeyMild respiratory disease and moderate egg decreaseWorldwidePRPSGRLa209484147
APMV3/Parakeet/Netherlands/449/1975Psittacines, passerines, flamingosNeurological, enteric, and respiratory diseaseWorldwideARPRGRLa171472314
APMV4/Duck/Hong Kong/D3/1975Duck, geese, chickensNone knownWorldwideVDIQPRF210076708
APMV5/Budgerigar/Japan/Kunitachi/1974Budgerigars, lorikeetsHigh mortality, enteric diseaseJapan, United Kingdom, AustraliaGKRKKRFa290563909
APMV6/Duck/Hong Kong/199/1977Ducks, geese, turkeysMild respiratory disease and increased mortality in turkeysWorldwidePAPEPRLb15081567
APMV7/Dove/TN/4/1975Pigeons, doves, turkeysMild respiratory disease in turkeysUnited States, England, JapanTLPSSRF224979458
APMV8/Goose/DE/1053/1976Ducks, geeseNone knownUnited States, JapanTYPQTRLa226343050
APMV9/Duck/NY/22/1978DucksNone knownWorldwideRIREGRIa217068693
APMV10/Penguin/Falkland Islands/324/2007Rockhopper penguinsNone KnownFalkland IslandsDKPSQRIa300432141
Open in a separate windowaRequires the addition of an exogenous protease.bProtease requirement depends on the isolate examined.cPutative.Six of these serotypes were classified in the latter half of the 1970s, when the most reliable assay available to classify paramyxoviruses was the hemagglutination inhibition (HI) assay (61). However, there are multiple problems associated with the use of serology, including the inability to classify some APMVs by comparing them to the sera of the nine defined APMVs alone (2, 8). In addition, one-way antigenicity and cross-reactivity between different serotypes have been documented for many years (4, 5, 14, 25, 29, 33, 34, 41, 51, 52, 60). The ability of APMVs, like other viruses, to show antigenic drift as it evolves over time (37, 43, 54) and the wide use and availability of precise molecular methods, such as PCR and genome sequencing, demonstrate the need for a more practical classification system.The genetic diversity of APMVs is still largely unexplored, as hundreds of avian species have never been surveyed for the presence of viruses that do not cause significant signs of disease or are not economically important. The emergence of H5N1 highly pathogenic avian influenza (HPAI) virus as the cause of the largest outbreak of a virulent virus in poultry in the past 100 years has spurred the development of surveillance programs to better understand the ecology of avian influenza (AI) viruses in aquatic birds around the globe, and in some instances it has provided opportunities for observing other viruses in wild bird populations (15, 53). In 2007, as part of a seabird health surveillance program in the Falkland Islands (Islas Malvinas), oral and cloacal swabs and serum were collected from rockhopper penguins (Eudyptes chrysocome) and environmental/fecal swab pools were collected from other seabirds.While AI virus has not yet been isolated from penguins in the sub-Antarctic and Antarctic areas, there have been two reports of serum antibodies positive to H7 and H10 from the Adélie species (11, 40). Rare isolations of APMV1, both virulent (45) and of low virulence (8), have been reported from Antarctic penguins. Sera positive for APMV1 and AMPV2 have also been reported (21, 24, 38, 40, 53). Since 1981, paramyxoviruses have been isolated from king penguins (Aptenodytes patagonicus), royal penguins (Eudyptes schlegeli), and Adélie penguins (Pygoscelis adeliae) from Antarctica and little blue penguins (Eudyptula minor) from Australia that cannot be identified as belonging to APMV1 to -9 and have not yet been classified (8, 11, 38-40). The morphology, biological and genomic characteristics, and antigenic relatedness of an APMV recently isolated from multiple penguin colonies on the Falkland Islands are reported here. Evidence that the virus belongs to a new serotype (APMV10) and a demonstration of the advantages of a whole genome system of analysis based on random sequencing followed by comparison of genetic distances are presented. Only after all APMVs are reported and classified will epidemiological information be known as to how the viruses are moving and spreading as the birds travel and interact with other avian species.  相似文献   

2.
3.
4.
5.
The human gut microbe Bacteroides fragilis can alter the expression of its surface molecules, such as capsular polysaccharides and SusC/SusD family outer membrane proteins, through reversible DNA inversions. We demonstrate here that DNA inversions at 12 invertible regions, including three gene clusters for SusC/SusD family proteins, were controlled by a single tyrosine site-specific recombinase (Tsr0667) encoded by BF0667 in B. fragilis strain YCH46. Genetic disruption of BF0667 diminished or attenuated shufflon-type DNA inversions at all three susC/susD genes clusters, as well as simple DNA inversions at nine other loci, most of which colocalized with susC/susD family genes. The inverted repeat sequences found within the Tsr0667-regulated invertible regions shared the consensus motif sequence AGTYYYN4GDACT. Tsr0667 specifically mediated the DNA inversions of 10 of the 12 regions, even under an Escherichia coli background when the invertible regions were exposed to BF0667 in E. coli cells. Thus, Tsr0667 is an additional globally acting DNA invertase in B. fragilis, which probably involves the selective expression of SusC/SusD family outer membrane proteins.The human gut harbors an abundant and diverse microbiota. Bacteroides is one of the most abundant genera of human gut microflora (10, 17, 20), and the biological activities of Bacteroides species are deeply integrated into human physiology through nutrient degradation, the production of short-chain fatty acids, or immunomodulatory molecules (11-14, 24). Recent genomic analyses of Bacteroides have revealed that the bacteria possess redundant abilities not only to bind and degrade otherwise indigestible dietary polysaccharides but also to produce vast arrays of capsular polysaccharide (5, 19, 38, 39). These functional redundancies have been established by the extensive duplication of various genes that encode molecules such as glycosylhydrolases, glycosyltransferases, and outer membrane proteins of the SusC/SusD family (starch utilization system) known to be involved in polysaccharide recognition and transport (7, 27, 28, 30). It has been assumed that these functional redundancies of Bacteroides contribute to the stability of the gut ecosystem (3, 21, 23, 32, 39).Another characteristic feature common in Bacteroides species is that the expression of some of the genes is altered in an on-off manner by reversible DNA inversions at gene promoters or within the protein-coding regions (5, 9, 19, 38, 39). These phase-variable phenotypes are associated with surface architectures such as capsular polysaccharides and SusC/SusD family proteins (5, 6, 16, 19). Our previous genomic analyses of Bacteroides fragilis strain YCH46 revealed that it contained as many as 31 invertible regions in its chromosome (19). These invertible regions can be grouped into six classes according to the internal motif sequences within inverted repeat sequences (IRs) (Table (Table1).1). The DNA inversions within these regions are thought to be controlled by site-specific DNA invertases specific to each class. B. fragilis strain YCH46 contains 33 tyrosine site-specific recombinases (Tsr) genes and four serine site-specific recombinases (Ssr) genes. Generally, DNA invertases mediate DNA inversions at adjacent regions, such as FimB and FimE, that flip their immediate downstream promoters to generate a phase-variable phenotype of type I pili in Escherichia coli (15). B. fragilis is unique in that this anaerobe possesses not only locally acting DNA invertases but also globally acting DNA invertases that mediate DNA inversions at distant loci (8, 29). It has been reported that B. fragilis possesses at least two types of master DNA invertase that regulate DNA inversions at multiple loci simultaneously (8, 29). One is Mpi, an Ssr that mediates the on-off switching of 13 promoter regions (corresponding to class I regions in B. fragilis strain YCH46), including seven promoter regions for capsular polysaccharide biosynthesis in B. fragilis strain NCTC9343 (8). The other master DNA invertase is Tsr19, a Tsr that regulates DNA inversions at two distantly located promoter regions (corresponding to class IV regions in B. fragilis strain YCH46) associated with the large encapsulation phenotype (6, 26, 29). The invertible regions contain specific consensus motifs within the IRs corresponding to each DNA invertase and constitute a regulatory unit. We designated this type of regulatory unit as an “inverlon,” which consists of at least two invertible regions controlled by a single master DNA invertase.

TABLE 1.

Classification of the invertible regions in B. fragilis strain YCH46 based on internal motif sequences within IRs
ClassaNo.Consensus motif sequencesbMaster DNA invertase genecRegulated genesSource or reference(s)
I14ARACGTWCGTBF2765 (mpi)Capsular polysaccharide biosynthesis genes8
II10AGTTC{N5}GAACTBF0667susC/susD paralogsThis study
III3GTTAC{N7}GTAACBF3038, BF4033, BF4283Putative outer membrane protein genes36
IV2TACTTANTAGGTAANAGAABF2766Extracellular polysacharide biosynthesis genes6, 26, 29
V1TCTGCAAAGNCTTTGCAGABF0667susC/susD paralogsThis study
VI1ACTAAGTTCTATCGGBF0667susC/susD paralogsThis study
Open in a separate windowaOur previous classification of the invertible regions identified in B. fragilis strain YCH46 genome (19).bConsensus motif sequences found within IRs are shown. R = A or G, W = A or T, and N = A, G, C, or T.cThe gene identifications in B. fragilis strain YCH46 genome are shown.Our previous studies indicated that an additional inverlon other than the Mpi- and Tsr19-regulated inverlons is present in B. fragilis, based on the finding that at least 10 invertible regions (corresponding to class II regions in B. fragilis strain YCH46) contain a particular consensus motif sequence (AAGTTCN5GAACTT) within their IRs (19) but do not appear to colocalize with a DNA invertase gene. The majority of the class II regions were associated with the selective switching of a particular set of susC/susD family genes. Since the SusC/SusD family of outer membrane proteins play an important role in polysaccharide utilization by Bacteroides (3, 23, 32), the inverlon associated with the phase variation of SusC/SusD family proteins would likely be involved in the survival of this anaerobe in the distal gut.In the present study, we sought to identify the DNA invertase regulating the additional inverlon in B. fragilis. Our results indicated that the Tsr encoded by BF0667 is a master DNA invertase for this inverlon (designated the Tsr0667-inverlon) in B. fragilis.  相似文献   

6.
Predator-prey relationships among prokaryotes have received little attention but are likely to be important determinants of the composition, structure, and dynamics of microbial communities. Many species of the soil-dwelling myxobacteria are predators of other microbes, but their predation range is poorly characterized. To better understand the predatory capabilities of myxobacteria in nature, we analyzed the predation performance of numerous Myxococcus isolates across 12 diverse species of bacteria. All predator isolates could utilize most potential prey species to effectively fuel colony expansion, although one species hindered predator swarming relative to a control treatment with no growth substrate. Predator strains varied significantly in their relative performance across prey types, but most variation in predatory performance was determined by prey type, with Gram-negative prey species supporting more Myxococcus growth than Gram-positive species. There was evidence for specialized predator performance in some predator-prey combinations. Such specialization may reduce resource competition among sympatric strains in natural habitats. The broad prey range of the Myxococcus genus coupled with its ubiquity in the soil suggests that myxobacteria are likely to have very important ecological and evolutionary effects on many species of soil prokaryotes.Predation plays a major role in shaping both the ecology and evolution of biological communities. The population and evolutionary dynamics of predators and their prey are often tightly coupled and can greatly influence the dynamics of other organisms as well (1). Predation has been invoked as a major cause of diversity in ecosystems (11, 12). For example, predators may mediate coexistence between superior and inferior competitors (2, 13), and differential trajectories of predator-prey coevolution can lead to divergence between separate populations (70).Predation has been investigated extensively in higher organisms but relatively little among prokaryotes. Predation between prokaryotes is one of the most ancient forms of predation (27), and it has been proposed that this process may have been the origin of eukaryotic cells (16). Prokaryotes are key players in primary biomass production (44) and global nutrient cycling (22), and predation of some prokaryotes by others is likely to significantly affect these processes. Most studies of predatory prokaryotes have focused on Bdellovibrionaceae species (e.g., see references 51, 55, and 67). These small deltaproteobacteria prey on other Gram-negative cells, using flagella to swim rapidly until they collide with a prey cell. After collision, the predator cells then enter the periplasmic space of the prey cell, consume the host cell from within, elongate, and divide into new cells that are released upon host cell lysis (41). Although often described as predatory, the Bdellovibrionaceae may also be considered to be parasitic, as they typically depend (apart from host-independent strains that have been observed [60]) on the infection and death of their host for their reproduction (47).In this study, we examined predation among the myxobacteria, which are also deltaproteobacteria but constitute a monophyletic clade divergent from the Bdellovibrionaceae (17). Myxobacteria are found in most terrestrial soils and in many aquatic environments as well (17, 53, 74). Many myxobacteria, including the model species Myxococcus xanthus, exhibit several complex social traits, including fruiting body formation and spore formation (14, 18, 34, 62, 71), cooperative swarming with two motility systems (64, 87), and group (or “wolf pack”) predation on both bacteria and fungi (4, 5, 8, 9, 15, 50). Using representatives of the genus Myxococcus, we tested for both intra- and interspecific variation in myxobacterial predatory performance across a broad range of prey types. Moreover, we examined whether prey vary substantially in the degree to which they support predatory growth by the myxobacteria and whether patterns of variation in predator performance are constant or variable across prey environments. The latter outcome may reflect adaptive specialization and help to maintain diversity in natural populations (57, 59).Although closely related to the Bdellovibrionaceae (both are deltaproteobacteria), myxobacteria employ a highly divergent mode of predation. Myxobacteria use gliding motility (64) to search the soil matrix for prey and produce a wide range of antibiotics and lytic compounds that kill and decompose prey cells and break down complex polymers, thereby releasing substrates for growth (66). Myxobacterial predation is cooperative both in its “searching” component (6, 31, 82; for details on cooperative swarming, see reference 64) and in its “handling” component (10, 29, 31, 32), in which secreted enzymes turn prey cells into consumable growth substrates (56, 83). There is evidence that M. xanthus employs chemotaxis-like genes in its attack on prey cells (5) and that predation is stimulated by close contact with prey cells (48).Recent studies have revealed great genetic and phenotypic diversity within natural populations of M. xanthus, on both global (79) and local (down to centimeter) scales (78). Phenotypic diversity includes variation in social compatibility (24, 81), the density and nutrient thresholds triggering development (33, 38), developmental timing (38), motility rates and patterns (80), and secondary metabolite production (40). Although natural populations are spatially structured and both genetic diversity and population differentiation decrease with spatial scale (79), substantial genetic diversity is present even among centimeter-scale isolates (78). No study has yet systematically investigated quantitative natural variation in myxobacterial predation phenotypes across a large number of predator genotypes.Given the previous discovery of large variation in all examined phenotypes, even among genetically extremely similar strains, we anticipated extensive predatory variation as well. Using a phylogenetically broad range of prey, we compared and contrasted the predatory performance of 16 natural M. xanthus isolates, sampled from global to local scales, as well as the commonly studied laboratory reference strain DK1622 and representatives of three additional Myxococcus species: M. flavescens (86), M. macrosporus (42), and M. virescens (63) (Table (Table1).1). In particular, we measured myxobacterial swarm expansion rates on prey lawns spread on buffered agar (31, 50) and on control plates with no nutrients or with prehydrolyzed growth substrate.

TABLE 1.

List of myxobacteria used, with geographical origin
Organism abbreviation used in textSpeciesStrainGeographic originReference(s)
A9Myxococcus xanthusA9Tübingen, Germany78
A23Myxococcus xanthusA23Tübingen, Germany78
A30Myxococcus xanthusA30Tübingen, Germany78
A41Myxococcus xanthusA41Tübingen, Germany78
A46Myxococcus xanthusA46Tübingen, Germany78
A47Myxococcus xanthusA47Tübingen, Germany78
A75Myxococcus xanthusA75Tübingen, Germany78
A85Myxococcus xanthusA85Tübingen, Germany78
TVMyxococcus xanthusTvärminneTvärminne, Finland79
PAKMyxococcus xanthusPaklenicaPaklenica, Croatia79
MADMyxococcus xanthusMadeira 1Madeira, Portugal79
WARMyxococcus xanthusWarwick 1Warwick, UK79
TORMyxococcus xanthusToronto 1Toronto, Ontario, Canada79
SUL2Myxococcus xanthusSulawesi 2Sulawesi, Indonesia79
KALMyxococcus xanthusKalalauKalalau, HI79
DAVMyxococcus xanthusDavis 1ADavis, CA79
GJV1Myxococcus xanthusGJV 1Unknown35, 72
MXFL1Myxococcus flavescensMx fl1Unknown65
MXV2Myxococcus virescensMx v2Unknown65
CCM8Myxococcus macrosporusCc m8Unknown65
Open in a separate window  相似文献   

7.
8.
9.
10.
11.
The inhibitor peptide DT-2 (YGRKKRRQRRRPPLRKKKKKH) is the most potent and selective inhibitor of the cGMP-dependent protein kinase (PKG) known today. DT-2 is a construct of a PKG tight binding sequence (W45, LRKKKKKH, KI = 0.8 μm) and a membrane translocating sequence (DT-6, YGRKKRRQRRRPP, KI = 1.1 μm), that combined strongly inhibits PKG catalyzed phosphorylation (KI = 12.5 nm) with ∼1000-fold selectivity toward PKG over protein kinase A, the closest relative of PKG. However, the molecular mechanism behind this inhibition is not entirely understood. Using a combination of photoaffinity labeling, stable isotope labeling, and mass spectrometry, we have located the binding sites of PKG-specific substrate and inhibitor peptides. Covalent linkage of a PKG-specific substrate analogue was localized in the catalytic core on residues 356–372, also known as the glycine-rich loop, essential for ATP binding. By analogy, the individual inhibitor peptides W45 and DT-6 were also found to cross-link near the glycine-rich loop, suggesting these are both substrate competitive inhibitors. A bifunctional photoreactive analogue of DT-2 was found to generate dimers of PKG. This cross-linking induced covalent PKG dimerization was not observed for an N-terminal deletion mutant of PKG, which lacks the dimerization domain. In addition, non-covalent mass spectrometry was used to determine binding stoichiometry and binding order of the inhibitor peptides. Dimeric PKG binds two W45 and DT-6 peptides, whereas only one DT-2 molecule was observed to bind to the dimeric PKG. Taken together, these findings imply that (i) the two individual components making up DT-2 are both targeted against the substrate-binding site and (ii) binding of a single DT-2 molecule inactivates both PKG monomers simultaneously, which is an indication that (iii) in cGMP-activated PKG the catalytic centers of both subunits may be in each other''s proximity.Among the superfamily of protein kinases the two cyclic nucleotide-regulated protein kinases, cAMP-dependent protein kinase and cGMP-dependent protein kinase, form a closely related subfamily of serine/threonine protein kinases (14). Both proteins share several structural elements, such as the N-terminal dimerization domain, an autoinhibition site, two in-tandem cyclic nucleotide-binding sites, and a highly conserved catalytic core (Fig. 1, A and B). Despite these similarities, these two enzymes display differences, which account for their unique properties. Whereas PKA2 is nearly ubiquitous, PKG is primarily found in the lung, cerebellum, and smooth muscles (5, 6). From a structural point of view these cyclic nucleotide-dependent protein kinases differ as well. The holoenzyme of PKA is a tetramer composed of two regulatory and two catalytic subunits. The catalytic subunits are non-covalently attached to the regulatory subunit dimer. Upon interaction with cAMP, the catalytic subunits dissociate from the holoenzyme and are free to catalyze heterophosphorylation (Fig. 1C). The mammalian type I PKGs are homodimeric cytosolic proteins containing two identical polypeptides of ∼76 kDa. Alternative mRNA splicing produces type Iα and type Iβ PKG, which are identical proteins apart from their first ∼100 N-terminal residues (7). Each PKG subunit is composed of a regulatory and a catalytic domain on a single polypeptide chain. Consequently, when cGMP activates PKG, the catalytic and regulatory components remain physically attached (Fig. 1D). Within the catalytic domain PKA and PKG share a strong primary sequence homology (8). Not surprisingly, these enzymes also exhibit overlapping substrate specificities, a feature that often interferes with efforts to elucidate their distinct biological pathways. Peptide substrates with a primary amino acid sequence motif RRX(S/T)X are in general recognized by both PKA and PKG (9). Besides this strong overlapping substrate specificity, several studies report on subtle differences in determinants that discriminate for PKA and PKG substrate specificity (1016). To specifically discriminate between PKG and PKA activity in biological assays a highly specific PKG peptide inhibitor was developed (17). This peptide, YGRKKRRQRRRPPLRKKKKKH (DT-2), is the most potent and selective PKG inhibitor known today. Recently, the validity of DT-2 as a superior inhibitor of PKG in terms of potency, selectivity, and membrane permeability has been demonstrated (1824). The inhibitor is a construct of a substrate competitive sequence, LRKKKKKH (W45), derived from a library screen that selected for tight PKG binding sequences, with a significant specificity toward PKG over PKA, and a membrane translocating signal peptide, YGRKKRRQRRRPP (DT-6). DT-2 strongly inhibits PKG-catalyzed phosphorylation (Ki = 12.5 nm), however, the molecular nature of DT-2 inhibition is not entirely understood (25). Because high resolution structural data are not available for PKG, one of our goals is to elucidate binding sites for PKG-specific substrates and inhibitors in more detail using a combination of mass spectrometric techniques and photoaffinity labeling. To further delineate the nature of inhibition we have developed photoaffinity analogues of DT-2 and related inhibitory peptides, as well as a high affinity peptide substrate. The method of photoaffinity labeling enables the direct probing of target proteins through a covalent bond, which is photochemically introduced between a ligand and its specific receptor (26). In combination with modern mass spectrometric techniques this is a powerful approach for the characterization of peptide-protein interactions (27). Substrate and inhibitor peptides containing photoactivatable analogues of phenylalanine, 4-benzoyl-l-phenylalanine (Phe(Bz)) or 4′-(3-(trifluoromethyl)-3H-diazirin-3-yl)-l-phenylalanine (Phe(Tmd)) were synthesized and used to locate their substrate/inhibitor-binding sites on PKG. These measurements indicate that the substrate peptide resides near the glycine-rich loop within the catalytic domain and that the inhibitor peptides are directed similarly toward this substrate-binding site, thereby acting as competitive inhibitors. In addition, nanoflow electrospray ionization time of flight mass spectrometry (ESI-TOF-MS) was performed to study the interaction between DT-2 and PKG in more detail. ESI-MS has proven to be a useful tool to analyze the non-covalent interaction of proteins with ligands, oligonucleotides, peptides, or other proteins (2831). Using this technique, important information on conformational changes (3235), measurement of relative dissociation constants (36, 37), and sequential binding order and cooperativity (38, 39) can be obtained. ESI-MS confirms that PKG is primarily a homodimer and is able to bind four cGMP molecules. Binding of DT-2 was strongly enhanced in the presence of cGMP. Surprising is the observation that only one DT-2 molecule binds to dimeric PKG. The information derived from these measurements allows for molecular modeling and structural refinements of the next generation of PKG-selective inhibitors.Open in a separate windowFIGURE 1.Linear arrangement of the functional domains of the regulatory and catalytic subunit of PKA (A) and PKG (B) type I and schematic representation of the current working models of the activation process of PKA (C) and PKG (D) type 1. Binding of cAMP to the PKA induces a conformational change that results in the dissociation of the catalytic subunits. Binding of cGMP to PKG also induces a conformational change, which exposes the catalytic domains, but both catalytic domains remain near each other via the N-terminal dimerization domain. (Images adapted from Scholten et al. (4).)

TABLE 1

Inhibition contants (KI) of PKA- or PKG-specific peptide inhibitors and the PKA/PKG specificity index
PeptideSequencePKGKiPKAKiSpecificity index (PKA/PKG)Ref.
μmμm
PKI(5–24)TTYDFIASGRTGRRNAIHD-NH21500.0030.0002(11)
WW21TQAKRKKALAMA-NH27.5750100(11)
W45LRKKKKKH0.82 ± 0.33559680(17)
DT-6YGRGGRRQRRRPP1.1 ± 0.2226 ± 423.6(17)
DT-2YGRKKRRQRRRPPLRKKKKKH0.0125 ± 0.00316.5 ± 3.81320(17)
Open in a separate window  相似文献   

12.
Herpes simplex virus 1 nucleocapsids bud through the inner nuclear membrane (INM) into the perinuclear space to obtain a primary viral envelope. This process requires a protein complex at the INM composed of the UL31 and UL34 gene products. While it is clear that the viral kinase encoded by the US3 gene regulates the localization of pUL31/pUL34 within the INM, the molecular mechanism by which this is accomplished remains enigmatic. Here, we have determined the following. (i) The N terminus of pUL31 is indispensable for the protein''s normal function and contains up to six serines that are phosphorylated by the US3 kinase during infection. (ii) Phosphorylation at these six serines was not essential for a productive infection but was required for optimal viral growth kinetics. (iii) In the presence of active US3 kinase, changing the serines to alanine caused the pUL31/pUL34 complex to aggregate at the nuclear rim and caused some virions to accumulate aberrantly in herniations of the nuclear membrane, much as in cells infected with a US3 kinase-dead mutant. (iv) The replacement of the six serines of pUL31 with glutamic acid largely restored the smooth distribution of pUL34/pUL31 at the nuclear membrane and precluded the accumulation of virions in herniations whether or not US3 kinase was active but also precluded the optimal primary envelopment of nucleocapsids. These observations indicate that the phosphorylation of pUL31 by pUS3 represents an important regulatory event in the virion egress pathway that can account for much of pUS3''s role in nuclear egress. The data also suggest that the dynamics of pUL31 phosphorylation modulate both the primary envelopment and the subsequent fusion of the nascent virion envelope with the outer nuclear membrane.The UL31 and UL34 proteins of herpes simplex virus 1 (HSV-1) form a complex that accumulates at the inner nuclear membrane (INM) of infected cells (26, 27). This complex is essential for the budding of nucleocapsids through the INM into the perinuclear space (26, 28). pUL34 is a type 2 integral membrane protein with a 247-amino-acid nucleoplasmic domain that binds pUL31 and holds the latter in close approximation to the INM (16, 19, 26, 31, 36, 37). Both proteins become incorporated into nascent virions, indicating that they directly or indirectly interact with nucleocapsids during the budding event (27). Interestingly, the coexpression of the pseudorabies virus homologs of HSV pUL31 and pUL34 are sufficient to induce budding from the INM in the absence of other viral proteins (13).The most prominent model of nuclear egress proposes that the step following primary envelopment involves the fusion of the perinuclear virion envelope with the outer nuclear membrane (ONM), allowing subsequent steps in which the deenveloped capsid engages budding sites in the Golgi or trans-Golgi network (20, 32). The US3 protein is a promiscuous kinase that phosphorylates pUL31, pUL34, and several other viral and cellular components (1, 2, 5, 11, 15, 21-23, 25). In the absence of pUS3 kinase activity, (i) virions accumulate within distensions of the perinuclear space that herniate into the nucleoplasm (14, 27, 29), (ii) the pUL31/pUL34 complex is mislocalized at the nuclear rim from a smooth pattern to discrete foci that accumulate adjacent to nuclear membrane herniations (12, 14, 27, 29), and (iii) the onset of infectious virus production is delayed (21, 29).Aberrant accumulations of perinuclear virions similar to those observed in cells infected with US3 kinase-dead viruses have been observed in cells infected with viruses lacking the capacity to produce glycoproteins H and B (gH and gB, respectively) (8). Because these proteins are required for fusion with the plasma membrane or endocytic vesicles during HSV entry (3, 4, 9, 10, 18, 30, 33), it has been proposed that the accumulation of perinuclear virions in the absence of gH and gB reflects a failure in the apparatus that normally mediates the fusion between the nascent virion envelope and the ONM (8). By extension of this hypothesis, pUS3 might act to trigger or otherwise regulate this perinuclear fusion event.The substrate(s) of the pUS3 kinase responsible for the altered localization of the pUL31/pUL34 complex and the aberrant accumulation of perinuclear virions were heretofore unknown. In one study to identify such a substrate, it was determined that precluding the phosphorylation of pUL34 was not responsible for the nuclear egress defects induced by the absence of pUS3 or its kinase activity (29). The current study was therefore undertaken to investigate the hypothesis that the pUS3-mediated phosphorylation of pUL31 is critical to regulate nuclear egress. The presented evidence indicates that aspects of the US3 kinase-dead phenotype, including the retention of virions in the perinuclear space, the mislocalization of the pUL31/pUL34 complex, and the delayed onset of virus replication, can be replicated by precluding pUL31 phosphorylation in the presence or absence of pUS3 kinase activity. The data also suggest that the dynamic phosphorylation of pUL31 is important during the primary envelopment of nucleocapsids.  相似文献   

13.
Degenerate primers for the [FeFe] hydrogenase (hydA) were developed and used in PCRs to examine hydA in microbial mats that inhabit saltern evaporative ponds in Guerrero Negro (GN), Mexico. A diversity of deduced HydA was discovered that revealed unique variants, which may reflect adaptation to the environmental conditions present in GN.Hydrogen (H2) is an important intermediate in the decomposition of organic matter in anaerobic environments and is the basis for many syntrophic interactions that occur in the food chain such as those between acetogens and methanogens and/or sulfate reducers. Little is known concerning the potential role of fermentative bacteria and/or H2 cycling in phototrophic microbial mat systems, such as the 6-cm-thick and extremely diverse microbial mats inhabiting the saline ponds at the Exportadora de Sal SA saltern in Guerrero Negro (GN), Baja California Sur, Mexico. Recent 16S rRNA gene surveys in GN revealed an abundance of bacteria in the upper 2 mm of the mat that, based on their phylogenetic affiliation, are thought to harbor fermentative metabolisms (6, 8, 13, 16, 20, 28). Similarly, previous studies have shown that the flux of H2 from the surface of the phototrophic mats present in GN is ∼150-fold higher during the night than during the day (9). Together, these results suggest the potential for fermentative metabolisms in H2 and carbon cycling in the GN saline mat environment.The [FeFe] hydrogenase has a central role in primary and secondary substrate fermentations, catalyzing the oxidation of excess reducing equivalents coupled to the reduction of protons, yielding H2 (2, 17, 29, 30). The [FeFe] hydrogenase is only known to occur in anaerobic bacteria and a small number of green algae (23, 29, 30), making this a useful biomarker for examining the diversity and distribution of this functional class of organism. Eighty-two putative [FeFe] hydrogenase large-subunit protein sequences (HydA) present in the GenBank database that represent the known diversity of putative HydA were screened for the L1 ([FLI]TSC[C/S]P[GAS]W[VIQH]) and L2 ([IVLF]MPCx[ASRD]K[KQ]xE) (conserved residues are in bold and underlined, and bracketed positions indicate “semiconserved” residues at that position) signature sequence motifs (17, 30). Degenerate PCR primers, corresponding to positions 272 to 279 (AD[M/L]TIMEE) and positions 420 to 427 (TGGVMEAA) in the Clostridium pasteurianum protein sequence, were designed for use in specific gene amplifications.Mat core samples (1 by 6 cm) were collected at 2:00 pm from pond 4 (35°C, pH 8.0, and 80 ppt salt) near pond 5 at the Exportadora de Sal saltworks, GN, on 13 February 2005. Mat cores were sectioned in 1-mm increments on site and immediately flash frozen in liquid nitrogen as previously described (16). Genomic DNA was extracted by bead beating in the presence of phenol-chloroform-isoamyl alcohol and sodium dodecyl sulfate as previously described (7) and was quantified by using the High DNA Mass Ladder (Invitrogen, Carlsbad, CA). Approximately 500-bp fragments of hydA were amplified in triplicate using primers FeFe-272F (5′-GCHGAYMTBACHATWATGGARGA-3′, where H = A, C, T; Y = C, T; M = A, C; B = C, G, T; W = A, T; R = A, G; 432-fold degeneracy) and FeFe-427R (5′-GCNGCYTCCATDACDCCDCCNGT-3′, where N = A, C, T, G; Y = C, T; D = A, G, T; 864-fold degeneracy) and 35 cycles of PCR as previously described (4). An empirically determined annealing temperature of 56.5°C, a MgCl2 concentration of 1.5 mM, and a primer concentration of 1 μM for each forward and reverse primer were utilized in 50-μl PCR mixtures containing 10 ng of genomic DNA as the template. Equal 40-μl volumes of three replicate PCR products were pooled, purified using the Promega Wizard purification kit (Madison, WI), quantified using the Low Mass DNA Ladder (Invitrogen), cloned using the pGEM Easy Vector System (Promega), and sequenced by using the M13F-M13R primer pair as previously described (5).Deduced amino acid sequences were screened for the presence of the L1 and L2 HydA sequence motifs as described above. ClustalX (version 2.0.8) (15) was employed to align inferred amino acid sequences and to create distance matrices for use in identifying and clustering operational taxonomic units (OTUs) with DOTUR (26). Calculations of Shannon diversity and Chao1 and Ace1 deduced amino acid sequence richness were completed with DOTUR by using the furthest-neighbor algorithm and a precision of 0.01. The phylogenetic position of putative HydA was assessed with MRBAYES (10, 25). Tree topologies were sampled every 500 generations for 2,000,000 generations (burnin = 1,000,000) by using the WAG evolutionary model with fixed amino acid frequencies and gamma-shaped rate variation with a proportion of invariable sites as recommended by ProtTest (1). The Saccharomyces cerevisiae Narf protein, a homolog of HydA, was used as the outgroup. The phylogenetic tree was projected using TreeView (version 1.6.6) (21).An unexpectedly diverse assemblage of putative HydA variants was present in the top 1 mm of the GN microbial mat, corresponding to the photic zone. At a sequence identity threshold (SIT) of 100%, the predicted Shannon diversity index was 3.84 and the mean Chao1 HydA richness was 187 unique OTUs (see Fig. S1a in the supplemental material). When a 99% SIT was applied, the predicted Chao1 richness estimate decreased 56% to 82 unique OTUs and the Shannon diversity index decreased slightly to 3.55 (data not shown). A further decrease in the SIT from 99 to 85% resulted in a slight decrease (23%) in the Chao1 predicted richness and a modest decrease in the Shannon diversity index from 3.55 to 3.28, suggesting that the majority of the putative HydA OTUs present in the top 1 mm of the mat is a result of recent evolution or speciation.The putative HydA variants from GN resolved into 42 OTUs which clustered into one of seven distinct phylogenetic clusters (Fig. (Fig.1A),1A), although clusters 5 and 7 could not be adequately resolved. Intriguingly, the mean identity scores determined from sequence alignment matrices for members of each of the seven sequence clusters recovered from GN, when compared to sequences in the GenBank database, were very low (47.5 to 73.2%) (Table (Table1),1), suggesting that the putative HydA sequences recovered from GN represent novel sequence space.Open in a separate windowFIG. 1.(A) Phylogenetic tree based on deduced putative HydA amino acid sequences from GN and reference HydA sequences. Nodes with posterior probabilities of <50% were collapsed in this analysis (posterior probabilities of 100 are denoted by asterisks). Bar, one substitution per 10 sites. (B) Partial-length deduced amino acid sequences of environmental and reference deduced HydA sequences illustrating the phylogenetic coherence of novel substitutions and insertions in and upstream of L1 sequence motif. Cluster designations (C1 to C7) correspond to those presented in Table Table11.

TABLE 1.

Phylogeny of partial-length putative HydA sequence clusters
Sequence clusteraNo. of clones in clusterMaximal intracluster sequence divergence (%)bMost closely related sequencecPhylumdMean % sequence identity (range)e
C12249.4Moorella thermoacetica ATCC 39073Firmicutes61.8 (56.4-67.3)
C21641.6Opitutus terrae PB90-1Verrucomicrobia64.3 (56.3-71.1)
C3628.8Bacteroides thetaiotaomicron VPI-5482Bacteroidetes72.1 (69.8-73.2)
C4938.6Alkaliphilus oremlandii OhILAsFirmicutes65.1 (61.4-66.4)
C5338.6Moorella thermoacetica ATCC 39073Firmicutes61.4 (NA)
C6751.8Heliobacillus mobilisFirmicutes50.3 (47.5-54.6)
C7232.9Thermoanaerobacterium saccharolyticum JW/SL-YS485Firmicutes61.2 (58.5-64.2)
Open in a separate windowaSequence clusters correspond to those presented in Fig. Fig.11.bMaximum intracluster sequence divergence among sequences within a cluster as calculated by the ClustalX sequence identity matrix following alignment using the Gonnet protein weight matrix (gap opening penalty of 13, gap extension penalty of 0.05).cThe most closely related sequence is defined as the taxon with the HydA sequence most closely related to that of members within the cluster as presented in Fig. Fig.11.dPhylum designation of most closely related HydA sequence as defined in Fig. Fig.11.eSequence identity means and ranges were calculated by using the ClustalX sequence identity matrix as described in footnote b. NA, not available.The highest proportion of putative HydA sequences recovered from the top 1-mm transect of the GN microbial mat were related to acetogens most closely affiliated with the Firmicutes (66.2% of the clones) and the Verrucomicrobia (24.6%), with a lower proportion related to the Bacteroidetes (9.2%). These observations support those of previous studies that identified the highest abundance of 16S rRNA genes related to fermentative/acetogenic organisms within the phyla Firmicutes, Verrucomicrobia, and Bacteroidetes in the upper few millimeters of microbial mat (16), coinciding with the mat transect where nighttime H2 concentrations were also elevated (9). The recovery of putative HydA sequences in the photic zone of the GN microbial mat is consistent with the results of previous studies conducted in the phototrophic mats in Yellowstone National Park, WY, which documented the potential for the secondary fermentation of cyanobacterial primary fermentation products in the top photic layers of the mat at night (3). Importantly, cyanobacterial fermentation may also contribute to the production of H2 in the mats during periods of anoxia (9).Deduced amino acid sequences from GN had significantly lower (P = <0.01 for both GRAVY and aliphatic indices) hydropathy indices than nonhalotolerant HydA sequences from GenBank (see Table S1 in the supplemental material), a feature which may reflect molecular adaptation to salinity. These observations are consistent with the results of previous investigations of deduced amino acid sequences from a variety of halophilic organisms that indicate hydrophobicity indices lower than those of their nonhalophilic counterparts, which is hypothesized to reduce the effects of salt-driven protein misfolding and/or aggregation (11, 12, 14, 24, 27, 31). In addition, many of the putative HydA sequences recovered from GN exhibited previously unobserved substitutions in the L1 motif and novel insertion domains upstream from the L1 motif (Fig. (Fig.1B).1B). Residues within L1 are involved in the coordination of the oxygen-labile [4Fe-4S] subcluster of the H cluster of HydA (18, 19, 22). Substitutions in this region may have implications for the redox properties of this [4Fe-4S] cluster and thus may be important in conferring stability in the presence of oxygen (23).The discovery of a rich assemblage of putative HydA variants in the top 1 mm of the GN microbial mat underscores the utility of these primers in examining the natural diversity of putative HydA and fermentative organisms in complex microbial ecosystems. The composition of putative HydA sequences from the top 1 mm of the microbial mat suggests adaption to the conditions present in the evaporative salterns in GN.  相似文献   

14.
15.
Ornithine decarboxylase (ODC) is the first enzyme involved in polyamine biosynthesis, and it catalyzes the decarboxylation of ornithine to putrescine. ODC is a dimeric enzyme, whereas antizyme inhibitor (AZI), a positive regulator of ODC that is homologous to ODC, exists predominantly as a monomer and lacks decarboxylase activity. The goal of this paper was to identify the essential amino acid residues that determine the dimerization of AZI. The nonconserved amino acid residues in the putative dimer interface of AZI (Ser-277, Ser-331, Glu-332, and Asp-389) were substituted with the corresponding residues in the putative dimer interface of ODC (Arg-277, Tyr-331, Asp-332, and Tyr-389, respectively). Analytical ultracentrifugation analysis was used to determine the size distribution of these AZI mutants. The size-distribution analysis data suggest that residue 331 may play a major role in the dimerization of AZI. Mutating Ser-331 to Tyr in AZI (AZI-S331Y) caused a shift from a monomer configuration to a dimer. Furthermore, in comparison with the single mutant AZI-S331Y, the AZI-S331Y/D389Y double mutant displayed a further reduction in the monomer-dimer Kd, suggesting that residue 389 is also crucial for AZI dimerization. Analysis of the triple mutant AZI-S331Y/D389Y/S277R showed that it formed a stable dimer (Kd value = 1.3 μm). Finally, a quadruple mutant, S331Y/D389Y/S277R/E332D, behaved as a dimer with a Kd value of ∼0.1 μm, which is very close to that of the human ODC enzyme. The quadruple mutant, although forming a dimer, could still be disrupted by antizyme (AZ), further forming a heterodimer, and it could rescue the AZ-inhibited ODC activity, suggesting that the AZ-binding ability of the AZI dimer was retained.Polyamines (putrescine, spermidine, and spermine) have been shown to have both structural and regulatory roles in protein and nucleic acid biosynthesis and function (13). Ornithine decarboxylase (ODC,3 EC 4.1.1.17) is a central regulator of cellular polyamine synthesis (reviewed in Refs. 1, 4, 5). This enzyme catalyzes the pyridoxal 5-phosphate (PLP)-dependent decarboxylation of ornithine to putrescine, and it is the first and rate-limiting enzyme in polyamine biosynthesis (2, 3, 6, 7). ODC and polyamines play important roles in a number of biological functions, including embryonic development, cell cycle, proliferation, differentiation, and apoptosis (815). They also have been associated with human diseases and a variety of cancers (1626). Because the regulation of ODC and polyamine content is critical to cell proliferation (11), as well as in the origin and progression of neoplastic diseases (23, 24), ODC has been identified as an oncogenic enzyme, and the inhibitors of ODC and the polyamine pathway are important targets for therapeutic intervention in many cancers (6, 11).ODC is ubiquitously found in organisms ranging from bacteria to humans. It contains 461 amino acid residues in each monomer and is a 106-kDa homodimer with molecular 2-fold symmetry (27, 28). Importantly, ODC activity requires the formation of a dimer (2931). X-ray structures of the ODC enzyme reveal that this dimer contains two active sites, both of which are formed at the interface between the N-terminal domain of one monomer, which provides residues involved in PLP interactions, and the C-terminal domain of the other subunit, which provides the residues that interact with substrate (27, 3241).ODC undergoes a unique ubiquitin-independent proteasomal degradation via a direct interaction with the regulatory protein antizyme (AZ). Binding of AZ promotes the dissociation of the ODC homodimers and targets ODC for degradation by the 26 S proteasome (4246). Current models of antizyme function indicate that increased polyamine levels promote the fidelity of the AZ mRNA translational frameshift, leading to increased concentrations of AZ (47). The AZ monomer selectively binds to dimeric ODC, thereby inactivating ODC by forming inactive AZ-ODC heterodimers (44, 4850). AZ acts as a regulator of polyamine metabolism that inhibits ODC activity and polyamine transport, thus restricting polyamine levels (4, 5, 51, 52). When antizymes are overexpressed, they inhibit ODC and promote ubiquitin-independent proteolytic degradation of ODC. Because elevated ODC activity is associated with most forms of human malignancies (1), it has been suggested that antizymes may function as tumor suppressors.In contrast to the extensive studies on the oncogene ODC, the endogenous antizyme inhibitor (AZI) is less well understood. AZI is homologous to the enzyme ODC. It is a 448-amino acid protein with a molecular mass of 50 kDa. However, despite the homology between these proteins, AZI does not possess any decarboxylase activity. It binds to antizyme more tightly than does ODC and releases ODC from the ODC-antizyme complex (53, 54). Both the AZI and AZ proteins display rapid ubiquitin-dependent turnover within a few minutes to 1 h in vivo (5). However, AZ binding actually stabilizes AZI by inhibiting its ubiquitination (55).AZI, which inactivates all members of the AZ family (53, 56), restores ODC activity (54), and prevents the proteolytic degradation of ODC, may play a role in tumor progression. It has been reported that down-regulation of AZI is associated with the inhibition of cell proliferation and reduced ODC activity, presumably through the modulation of AZ function (57). Moreover, overexpression of AZI has been shown to increase cell proliferation and promote cell transformation (5860). Furthermore, AZI is capable of direct interaction with cyclin D1, preventing its degradation, and this effect is at least partially independent of AZ function (60, 61). These results demonstrate a role for AZI in the positive regulation of cell proliferation and tumorigenesis.It is now known that ODC exists as a dimer and that AZI may exist as a monomer physiologically (62). Fig. 1 shows the dimeric structures of ODC (Fig. 1A) and AZI (Fig. 1B). Although structural studies indicate that both ODC and AZI crystallize as dimers, the dimeric AZI structure has fewer interactions at the dimer interface, a smaller buried surface area, and a lack of symmetry of the interactions between residues from the two monomers, suggesting that the AZI dimer may be nonphysiological (62). In this study, we identify the critical amino acid residues governing the difference in dimer formation between ODC and AZI. Our preliminary studies using analytical ultracentrifugation indicated that ODC exists as a dimer, whereas AZI exists in a concentration-dependent monomer-dimer equilibrium. Multiple sequence alignments of ODC and AZI from various species have shown that residues 277, 331, 332, and 389 are not conserved between ODC and AZI (Open in a separate windowFIGURE 1.Crystal structure and the amino acid residues at the dimer interface of human ornithine decarboxylase (hODC) and mouse antizyme inhibitor (mAZI). A, homodimeric structure of human ODC with the cofactor PLP analog, LLP (Protein Data Bank code 1D7K). B, putative dimeric structure of mouse AZI (Protein Data Bank code 3BTN). The amino acid residues in the dimer interface are shown as a ball-and-stick model. The putative AZ-binding site is colored in cyan. This figure was generated using PyMOL (DeLano Scientific LLC, San Carlos, CA).

TABLE 1

Amino acid residues at the dimer interface of human ODC and AZI
Human ODCResidueHuman AZI
Nonconserved
    Arg277Ser
    Tyr331Ser
    Asp332Glu
    Tyr389Asp

Conserved
    Asp134Asp
    Lys169Lys
    Lys294Lys
    Tyr323Tyr
    Asp364Asp
    Gly387Gly
    Phe397Phe
Open in a separate window  相似文献   

16.
Glycopeptidolipids (GPLs) are one of the major glycolipid components present on the surface of Mycobacterium avium complex (MAC) that belong to opportunistic pathogens distributed in the natural environment. The serovars of MAC, up to around 30 types, are defined by the variable oligosaccharide portions of the GPLs. Epidemiological studies show that serovar 4 is the most prevalent type, and the prognosis of pulmonary disease caused by serovar 4 is significantly worse than that caused by other serovars. However, little is known about the biosynthesis of serovar 4-specific GPL, particularly the formation of the oligosaccharide portion that determines the properties of serovar 4. To investigate the biosynthesis of serovar 4-specific GPL, we focused on one segment that included functionally unknown genes in the GPL biosynthetic gene cluster of a serovar 4 strain. In this segment, a putative hemolytic protein gene, hlpA, and its downstream gene were found to be responsible for the formation of the 4-O-methyl-rhamnose residue, which is unique to serovar 4-specific GPL. Moreover, functional characterization of the hlpA gene revealed that it encodes a rhamnosyltransferase that transfers a rhamnose residue via 1→4 linkage to a fucose residue of serovar 2-specific GPL, which is a key pathway leading to the synthesis of oligosaccharide of serovar 4-specific GPL. These findings may provide clues to understanding the biological role of serovar 4-specific GPL in MAC pathogenicity and may also provide new insights into glycosyltransferase, which generates structural and functional diversity of GPLs.The genus Mycobacterium has a unique feature in the cell envelope that contains a multilayered structure consisting of peptidoglycan, mycolyl-arabinogalactan complex, and surface glycolipids (8, 12). It is known that these components play a role in protection from environmental stresses, such as antimicrobial agents and host immune responses (8, 12). Some of them are recognized as pathogenic factors related to mycobacterial diseases, such as tuberculosis and leprosy (8, 12). In case of nontuberculous mycobacteria that are widely distributed in the natural environment as opportunistic pathogens, glycopeptidolipids (GPLs) are abundantly present on the cell envelope as surface glycolipids (34). GPLs have a core structure in which a fatty acyl-tetrapeptide is glycosylated with 6-deoxy-talose (6-d-Tal) and O-methyl-rhamnose (O-Me-Rha) (2, 5, 13). This structure is common to all types of GPLs, and GPLs with this structure that have not undergone further glycosylation are termed non-serovar-specific GPLs (nsGPLs) (2, 5, 13). Structural diversity generated by further glycosylations, such as rhamnosylation, fucosylation, and glucosylation, is observed for the oligosaccharide portion linked to the 6-d-Tal residue of nsGPLs from Mycobacterium avium complex (MAC), a member of the nontuberculous mycobacteria consisting of two species, M. avium and M. intracellulare (2, 5, 34). Consequently, these nsGPLs with varied oligosaccharides lead to the formation of the serovar-specific GPLs (ssGPLs) that define around 30 types of MAC serovars (10).The properties of MAC serovars are known to be notably different from each other and also to be closely associated with the pathogenicity of MAC (3, 6, 18, 30, 31, 32). Various epidemiological studies indicate that serovar 4 is the most prevalent type and is also one of the serovars frequently isolated from AIDS patients (1, 20, 33, 36). Additionally, pulmonary MAC disease caused by serovar 4 is shown to exhibit a poorer prognosis than that caused by other serovars (23). With respect to host immune responses to MAC infection, serovar 4-specific GPL is reported to have characteristic features that are in contrast to those of other ssGPLs (21, 30). Structurally, serovar 4-specific GPL contains a unique oligosaccharide in which the oligosaccharide of serovar 2-specific GPL is further glycosylated with 4-O-methyl-rhamnose (4-O-Me-Rha) residue through a 1→4 linkage (Table (Table1)1) (24). Therefore, it is thought that the presence of 4-O-Me-Rha and its linkage position are important in exhibiting the specificity of biological activities. The biosynthesis of the oligosaccharide portion in several ssGPLs is currently being clarified (15, 16, 17, 25, 26), while that of serovar 4-specific GPL is still unresolved. In this study, we have focused on the genomic region predicted to be associated with GPL biosynthesis in the serovar 4 strain and explored the key genes responsible for the formation of 4-O-Me-Rha that might determine the specific properties of MAC serovar 4.

TABLE 1.

Oligosaccharide structures of serovar 2- and 4-specific GPLs
SerovarOligosaccharideReference
22,3-di-O-Me-α-l-Fuc-(1→3)-α-l-Rha-(1→2)-l-6-d-Tal9
44-O-Me-α-l-Rha-(1→4)-2-O-Me-α-l-Fuc-(1→3)-α-l-Rha-(1→2)-l-6-d-Tal24
Open in a separate window  相似文献   

17.
UL31 and UL34 of herpes simplex virus type 1 form a complex necessary for nucleocapsid budding at the inner nuclear membrane (INM). Previous examination by immunogold electron microscopy and electron tomography showed that pUL31, pUL34, and glycoproteins D and M are recruited to perinuclear virions and densely staining regions of the INM where nucleocapsids bud into the perinuclear space. We now show by quantitative immunogold electron microscopy coupled with analysis of variance that gD-specific immunoreactivity is significantly reduced at both the INM and outer nuclear membrane (ONM) of cells infected with a UL34 null virus. While the amount of gM associated with the nuclear membrane (NM) was only slightly (P = 0.027) reduced in cells infected with the UL34 null virus, enrichment of gM in the INM at the expense of that in the ONM was greatly dependent on UL34 (P < 0.0001). pUL34 also interacted directly or indirectly with immature forms of gD (species expected to reside in the endoplasmic reticulum or nuclear membrane) in lysates of infected cells and with the cytosolic tail of gD fused to glutathione S-transferase in rabbit reticulocyte lysates, suggesting a role for the pUL34/gD interaction in recruiting gD to the NM. The effects of UL34 on gD and gM localization were not a consequence of decreased total expression of gD and gM, as determined by flow cytometry. Separately, pUL31 was dispensable for targeting gD and gM to the two leaflets of the NM but was required for (i) the proper INM-versus-ONM ratio of gD and gM in infected cells and (ii) the presence of electron-dense regions in the INM, representing nucleocapsid budding sites. We conclude that in addition to their roles in nucleocapsid envelopment and lamina alteration, UL31 and UL34 play separate but related roles in recruiting appropriate components to nucleocapsid budding sites at the INM.Herpesvirus virions comprise a nucleocapsid containing genomic viral DNA, a proteinaceous tegument layer surrounding the nucleocapsid, and a virion envelope surrounding the tegument. The envelope of extracellular herpes simplex virus (HSV) virions contains glycoproteins gB, gC, gD, gE, gI, gG, gH, gK, gL, and gM (23, 51).As viewed by electron microscopy, nascent virions form as the nucleocapsid buds through densely staining regions of the nuclear membrane (NM) (21, 41). Electron tomograms of HSV perinuclear virions compared to those of extracellular virions infer that the former contain glycoproteins of considerably less glycosylation and a relatively sparse tegument layer compared to their counterparts in mature extracellular virions (6). The lower levels of glycosylation in HSV perinuclear virions are consistent with the fact that the lumen of the perinuclear space is continuous with that of the endoplasmic reticulum. Thus, the polysaccharide moieties of virion glycoproteins become fully processed as virions access Golgi enzymes during their egress to the extracellular space. Although the full proteome of the nascent perinuclear virion is unknown, immunogold studies have shown that they contain at least pUL31, pUL34, pUS3, gB, gC, gD, gH, gM, and the VP16 and pUL11 tegument proteins in addition to the proteins that comprise the viral capsid (4, 5, 15, 25, 37, 40, 47, 50, 55).The UL31 and UL34 gene products of HSV-1 (pUL31 and pUL34, respectively) form a complex that localizes at the inner and outer NMs (INM and ONM, respectively) of infected cells (40). Both proteins are essential for nucleocapsid envelopment at the INM and become incorporated into nascent virions when nucleocapsids bud through the INM into the perinuclear space (39, 40, 42). The proteins and their essential role in nucleocapsid envelopment are conserved in all herpesvirus subfamilies (14, 20, 32, 45). pUL31 of HSV-1 is a mostly hydrophobic phosphoprotein that is held in close approximation to the nucleoplasmic face of the INM by interaction with pUL34, an integral membrane protein of type II orientation (33, 40, 46, 56). The first 248 amino acids of pUL34 are predicted to reside in the nucleoplasm or cytoplasm, depending on whether the protein localizes in the INM or ONM, respectively. This is followed by an approximately 22-amino acid transmembrane domain with up to 5 amino acids residing in the perinuclear space or lumen of the endoplasmic reticulum.In the most prominent model of herpesvirion egress, the envelope of the perinuclear virion fuses with the ONM, releasing the deenveloped nucleocapsid into the cytoplasm, where it subsequently buds into cytoplasmic membranous organelles such as the Golgi or trans-Golgi network (34, 49). This model is supported by the observation that pUL31 and pUL34 are located in the perinuclear virion but not extracellular virions (18, 40). Thus, these proteins are lost from the virion upon fusion of the virion envelope with the ONM. Also supporting this egress model is the observation that deletion of both gB and gH causes virions to accumulate aberrantly in the perinuclear space (15). The involvement of gH and gB is potentially satisfying because these proteins comprise essential components of the machinery that mediates fusion of the virion envelope with the plasma or endosomal membranes during the initiation of infection (9, 12, 16, 44, 52). Moreover, expression of a combination of gB, gD, gH, and gL is sufficient to mediate fusion of cell membranes, whereas coexpression with gM or gK inhibits this fusion (3, 8, 11). Although the mechanism of fusion is unclear, gD is known to bind viral receptors on cell surfaces, and the structure of gB indicates features reminiscent of other viral fusion proteins (24, 35, 48). gD has been shown to interact with gB and gH at least transiently, suggesting that these interactions may be important for the fusion reaction (1, 2). Thus, fusion between the nascent and mature virion envelopes with target membranes may share mechanistic similarities.On the other hand, it is likely that the two fusion events are mechanistically distinct because (i) single deletion of either gH or gB precludes viral entry and cell/cell fusion but does not cause nascent virions to accumulate in the perinuclear space (9, 16, 31, 43) and (ii) the activity of a viral kinase encoded by US3 is dispensable for entry but believed to promote fusion of the perinuclear virion and ONM (28, 40). Moreover, the lack of glycoproteins from the pseudorabies virus perinuclear virion suggests that fusion is mediated by an entirely different mechanism in this system (26).The current study focuses on how glycoproteins are incorporated into the nascent virion. We show that optimal recruitment of gD to both leaflets of the NM and gM to the INM requires pUL34 and pUL31. We also show that immature gD interacts with pUL34, suggesting a mechanism by which pUL34 might recruit gD to the NM.  相似文献   

18.
19.
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