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.
Andes virus (ANDV) causes a fatal hantavirus pulmonary syndrome (HPS) in humans and Syrian hamsters. Human αvβ3 integrins are receptors for several pathogenic hantaviruses, and the function of αvβ3 integrins on endothelial cells suggests a role for αvβ3 in hantavirus directed vascular permeability. We determined here that ANDV infection of human endothelial cells or Syrian hamster-derived BHK-21 cells was selectively inhibited by the high-affinity αvβ3 integrin ligand vitronectin and by antibodies to αvβ3 integrins. Further, antibodies to the β3 integrin PSI domain, as well as PSI domain polypeptides derived from human and Syrian hamster β3 subunits, but not murine or bovine β3, inhibited ANDV infection of both BHK-21 and human endothelial cells. These findings suggest that ANDV interacts with β3 subunits through PSI domain residues conserved in both Syrian hamster and human β3 integrins. Sequencing the Syrian hamster β3 integrin PSI domain revealed eight differences between Syrian hamster and human β3 integrins. Analysis of residues within the PSI domains of human, Syrian hamster, murine, and bovine β3 integrins identified unique proline substitutions at residues 32 and 33 of murine and bovine PSI domains that could determine ANDV recognition. Mutagenizing the human β3 PSI domain to contain the L33P substitution present in bovine β3 integrin abolished the ability of the PSI domain to inhibit ANDV infectivity. Conversely, mutagenizing either the bovine PSI domain, P33L, or the murine PSI domain, S32P, to the residue present human β3 permitted PSI mutants to inhibit ANDV infection. Similarly, CHO cells transfected with the full-length bovine β3 integrin containing the P33L mutation permitted infection by ANDV. These findings indicate that human and Syrian hamster αvβ3 integrins are key receptors for ANDV and that specific residues within the β3 integrin PSI domain are required for ANDV infection. Since L33P is a naturally occurring human β3 polymorphism, these findings further suggest the importance of specific β3 integrin residues in hantavirus infection. These findings rationalize determining the role of β3 integrins in hantavirus pathogenesis in the Syrian hamster model.Hantaviruses persistently infect specific small mammal hosts and are spread to humans by the inhalation of aerosolized excreted virus (41, 42). Hantaviruses predominantly infect endothelial cells and cause one of two vascular leak-based diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (41). Hantavirus diseases are characterized by increased vascular permeability and acute thrombocytopenia in the absence of endothelial cell lysis (36, 41, 42, 54). In general, hantaviruses are not spread from person to person; however, the Andes hantavirus (ANDV) is an exception, since there are several reports of person-to-person transmission of ANDV infection (11, 37, 47, 52). ANDV is also unique in its ability to cause an HPS-like disease in Syrian hamsters and serves as the best-characterized hantavirus disease model with a long onset, symptoms, and pathogenesis nearly identical to that of HPS patients (20, 21, 50).Hantavirus infection of the endothelium alters endothelial cell barrier functions through direct and immunological responses (8, 14). Although the means by which hantaviruses cause pulmonary edema or hemorrhagic disease has been widely conjectured, the mechanisms by which hantaviruses elicit pathogenic human responses have yet to be defined. Hantaviruses coat the surface of infected VeroE6 cells days after infection (17), and this further suggests that dynamic hantavirus interactions with immune and endothelial cells are likely to contribute to viral pathogenesis. Hantavirus pathogenesis has been suggested to involve CD8+ T cells, tumor necrosis factor alpha or other cytokines, viremia, and the dysregulation of β3 integrins (7, 8, 13-16, 25-28, 32, 34, 38, 44-46). However, these responses have not been demonstrated to contribute to hantavirus pathogenesis, and in some cases there are conflicting data on their involvement (18, 25-28, 34, 35, 44, 45, 48). Immune complex deposition clearly contributes to HFRS patient disease and renal sequelae (4, 7), but it is unclear what triggers vascular permeability in HPS and HFRS diseases or why hemorrhage occurs in HFRS patients but not in HPS patients (8, 36, 54). Acute thrombocytopenia is common to both diseases, and platelet dysfunction resulting from defective platelet aggregation is reported in HFRS patients (7, 8).Pathogenic hantaviruses have in common their ability to interact with αIIbβ3 and αvβ3 integrins present on platelets and endothelial cells (13, 16), and β3 integrins have primary roles in regulating vascular integrity (1, 2, 6, 19, 22, 39, 40). Consistent with the presence of cell surface displayed virus (17), pathogenic hantaviruses uniquely block αvβ3 directed endothelial cell migration and enhance endothelial cell permeability for 3 to 5 days postinfection (14, 15). Pathogenic hantaviruses dysregulate β3 integrin functions by binding domains present at the apex of inactive β3 integrin conformers (38). αvβ3 forms a complex with vascular endothelial cell growth factor receptor 2 (VEGFR2) and normally regulates VEGF-directed endothelial cell permeability (2, 3, 10, 39, 40). However, both β3 integrin knockouts and hantavirus-infected endothelial cells result in increased VEGF-induced permeability, presumably by disrupting VEGFR2-β3 integrin complex formation (2, 14, 19, 39, 40). This suggests that at least one means for hantaviruses to increase vascular permeability occurs through interactions with β3 integrins that are required for normal platelet and endothelial cell functions.αvβ3 and αIIbβ3 integrins exist in two conformations: an active extended conformation where the ligand binding head domain is present at the apex of the heterodimer and a basal, inactive bent conformation where the globular head of the integrin is folded toward the cell membrane (30, 53, 55). Pathogenic HTN and NY-1 hantaviruses bind to the N-terminal plexin-semaphorin-integrin (PSI) domain of β3 integrin subunits and are selective for bent, inactive αvβ3 integrin conformers (38). Pathogenic hantavirus binding to inactive αvβ3 integrins is consistent with the selective inhibitory effect of hantaviruses on αvβ3 function and endothelial cell permeability (14, 15, 38). Although the mechanism of hantavirus induced vascular permeability has yet to be defined, there is a clear role for β3 integrin dysfunction in vascular permeability deficits (5, 6, 22, 29, 39, 40, 51) which make an understanding of hantavirus interactions with β3 subunits important for both entry and disease processes.The similarity between HPS disease in humans and Syrian hamsters (20, 21) suggests that pathogenic mechanisms of ANDV disease are likely to be coincident. Curiously, other hantaviruses (Sin Nombre virus [SNV] and Hantaan virus [HTNV]) are restricted in Syrian hamsters and fail to cause disease in this animal, even though they are prominent causes of human disease (50). Although the host range restriction for SNV and HTNV in Syrian hamsters has not been defined (33), the pathogenesis of ANDV in Syrian hamsters suggests that both human and Syrian hamster β3 integrins may similarly be used by ANDV and contribute to pathogenesis.We demonstrate here that ANDV infection of the Syrian hamster BHK-21 cell line and human endothelial cells is dependent on αvβ3 and inhibited by αvβ3 specific ligands and antibodies. Further, polypeptides expressing the N-terminal 53 residues of human and Syrian hamster β3 subunits block ANDV infection. This further indicates that ANDV interaction with the N-terminal 53 residues of both human and Syrian hamster β3 integrins is required for viral entry. We also demonstrate that ANDV recognition of human and Syrian hamster β3 integrins is determined by proline substitutions at residues 32/33 within the β3 integrin PSI domain. These results define unique ANDV interactions with human and Syrian hamster β3 integrins.  相似文献   

2.
3.
Spores of Bacillus subtilis contain a number of small, acid-soluble spore proteins (SASP) which comprise up to 20% of total spore core protein. The multiple α/β-type SASP have been shown to confer resistance to UV radiation, heat, peroxides, and other sporicidal treatments. In this study, SASP-defective mutants of B. subtilis and spores deficient in dacB, a mutation leading to an increased core water content, were used to study the relative contributions of SASP and increased core water content to spore resistance to germicidal 254-nm and simulated environmental UV exposure (280 to 400 nm, 290 to 400 nm, and 320 to 400 nm). Spores of strains carrying mutations in sspA, sspB, and both sspA and sspB (lacking the major SASP-α and/or SASP-β) were significantly more sensitive to 254-nm and all polychromatic UV exposures, whereas the UV resistance of spores of the sspE strain (lacking SASP-γ) was essentially identical to that of the wild type. Spores of the dacB-defective strain were as resistant to 254-nm UV-C radiation as wild-type spores. However, spores of the dacB strain were significantly more sensitive than wild-type spores to environmental UV treatments of >280 nm. Air-dried spores of the dacB mutant strain had a significantly higher water content than air-dried wild-type spores. Our results indicate that α/β-type SASP and decreased spore core water content play an essential role in spore resistance to environmentally relevant UV wavelengths whereas SASP-γ does not.Spores of Bacillus spp. are highly resistant to inactivation by different physical stresses, such as toxic chemicals and biocidal agents, desiccation, pressure and temperature extremes, and high fluences of UV or ionizing radiation (reviewed in references 33, 34, and 48). Under stressful environmental conditions, cells of Bacillus spp. produce endospores that can stay dormant for extended periods. The reason for the high resistance of bacterial spores to environmental extremes lies in the structure of the spore. Spores possess thick layers of highly cross-linked coat proteins, a modified peptidoglycan spore cortex, a low core water content, and abundant intracellular constituents, such as the calcium chelate of dipicolinic acid and α/β-type small, acid-soluble spore proteins (α/β-type SASP), the last two of which protect spore DNA (6, 42, 46, 48, 52). DNA damage accumulated during spore dormancy is also efficiently repaired during spore germination (33, 47, 48). UV-induced DNA photoproducts are repaired by spore photoproduct lyase and nucleotide excision repair, DNA double-strand breaks (DSB) by nonhomologous end joining, and oxidative stress-induced apurinic/apyrimidinic (AP) sites by AP endonucleases and base excision repair (15, 26-29, 34, 43, 53, 57).Monochromatic 254-nm UV radiation has been used as an efficient and cost-effective means of disinfecting surfaces, building air, and drinking water supplies (31). Commonly used test organisms for inactivation studies are bacterial spores, usually spores of Bacillus subtilis, due to their high degree of resistance to various sporicidal treatments, reproducible inactivation response, and safety (1, 8, 19, 31, 48). Depending on the Bacillus species analyzed, spores are 10 to 50 times more resistant than growing cells to 254-nm UV radiation. In addition, most of the laboratory studies of spore inactivation and radiation biology have been performed using monochromatic 254-nm UV radiation (33, 34). Although 254-nm UV-C radiation is a convenient germicidal treatment and relevant to disinfection procedures, results obtained by using 254-nm UV-C are not truly representative of results obtained using UV wavelengths that endospores encounter in their natural environments (34, 42, 50, 51, 59). However, sunlight reaching the Earth''s surface is not monochromatic 254-nm radiation but a mixture of UV, visible, and infrared radiation, with the UV portion spanning approximately 290 to 400 nm (33, 34, 36). Thus, our knowledge of spore UV resistance has been constructed largely using a wavelength of UV radiation not normally reaching the Earth''s surface, even though ample evidence exists that both DNA photochemistry and microbial responses to UV are strongly wavelength dependent (2, 30, 33, 36).Of recent interest in our laboratories has been the exploration of factors that confer on B. subtilis spores resistance to environmentally relevant extreme conditions, particularly solar UV radiation and extreme desiccation (23, 28, 30, 34 36, 48, 52). It has been reported that α/β-type SASP but not SASP-γ play a major role in spore resistance to 254-nm UV-C radiation (20, 21) and to wet heat, dry heat, and oxidizing agents (48). In contrast, increased spore water content was reported to affect B. subtilis spore resistance to moist heat and hydrogen peroxide but not to 254-nm UV-C (12, 40, 48). However, the possible roles of SASP-α, -β, and -γ and core water content in spore resistance to environmentally relevant solar UV wavelengths have not been explored. Therefore, in this study, we have used B. subtilis strains carrying mutations in the sspA, sspB, sspE, sspA and sspB, or dacB gene to investigate the contributions of SASP and increased core water content to the resistance of B. subtilis spores to 254-nm UV-C and environmentally relevant polychromatic UV radiation encountered on Earth''s surface.  相似文献   

4.
5.
6.
The β clamp is an essential replication sliding clamp required for processive DNA synthesis. The β clamp is also critical for several additional aspects of DNA metabolism, including DNA mismatch repair (MMR). The dnaN5 allele of Bacillus subtilis encodes a mutant form of β clamp containing the G73R substitution. Cells with the dnaN5 allele are temperature sensitive for growth due to a defect in DNA replication at 49°C, and they show an increase in mutation frequency caused by a partial defect in MMR at permissive temperatures. We selected for intragenic suppressors of dnaN5 that rescued viability at 49°C to determine if the DNA replication defect could be separated from the MMR defect. We isolated three intragenic suppressors of dnaN5 that restored growth at the nonpermissive temperature while maintaining an increase in mutation frequency. All three dnaN alleles encoded the G73R substitution along with one of three novel missense mutations. The missense mutations isolated were S22P, S181G, and E346K. Of these, S181G and E346K are located near the hydrophobic cleft of the β clamp, a common site occupied by proteins that bind the β clamp. Using several methods, we show that the increase in mutation frequency resulting from each dnaN allele is linked to a defect in MMR. Moreover, we found that S181G and E346K allowed growth at elevated temperatures and did not have an appreciable effect on mutation frequency when separated from G73R. Thus, we found that specific residue changes in the B. subtilis β clamp separate the role of the β clamp in DNA replication from its role in MMR.Replication sliding clamps are essential cellular proteins imparting a spectacular degree of processivity to DNA polymerases during genome replication (24, 39-41). Encoded by the dnaN gene, the β clamp is a highly conserved bacterial sliding clamp found in virtually all eubacterial species (reviewed in reference 7). The β clamp is a head-to-tail, ring-shaped homodimer that encircles double-stranded DNA (1, 39). In eukaryotes and archaea, the analog of the β clamp is proliferating cell nuclear antigen (PCNA) (15, 28, 40, 41). Eukaryotic PCNA is a ring-shaped homotrimer that also acts to encircle DNA, increasing the processivity of the replicative DNA polymerases (40, 41). Although the primary structures of the β clamp and PCNA are not conserved, the tertiary structures of these proteins are very similar, demonstrating structural conservation among bacterial, archaeal, and eukaryotic replication sliding clamps (28, 39-41; reviewed in reference 6).The function of the β clamp is not limited to its well-defined role in genome replication. The Escherichia coli β clamp binds Hda, which also binds the replication initiation protein DnaA, regulating the active form of DnaA complexed with ATP (19, 37, 43). This allows the β clamp to regulate replication initiation through the amount of available DnaA-ATP. In Bacillus subtilis, the β clamp binds YabA, a negative regulator of DNA replication initiation (12, 29, 52). It has also been suggested that the B. subtilis β clamp sequesters DnaA from the replication origin during the cell cycle through the binding of DnaA to YabA and the binding of YabA to the β clamp (70). Thus, it is hypothesized that in E. coli and B. subtilis, the β clamp influences the frequency of replication initiation through interactions with Hda and YabA, respectively.The E. coli and B. subtilis β clamp has an important role in translesion DNA synthesis during the replicative bypass of noncoding bases by specialized DNA polymerases belonging to the Y family (20, 33). The roles of the E. coli β clamp in translesion synthesis are well established (5, 8, 30, 31). Binding sites on the E. coli β clamp that accommodate translesion polymerases pol IV (DinB) and pol V (UmuD2′C) have been identified, and the consequence of disrupting their association with the β clamp has illustrated the critical importance of the β clamp to the activity of both of these polymerases (4, 5, 8, 26, 30, 31, 48, 49).In addition to the involvement of the β clamp in replication initiation, DNA replication, and translesion synthesis, the E. coli and B. subtilis β clamp also functions in DNA mismatch repair (MMR) (45, 46, 64). The MMR pathway recognizes and repairs DNA polymerase errors, contributing to the overall fidelity of the DNA replication pathway (reviewed in references 42 and 60). In both E. coli and B. subtilis, deletion of the genes mutS and mutL increases the spontaneous mutation frequency several hundredfold (13, 25, 63). In E. coli, MutS recognizes and binds mismatches, while MutL functions as a “matchmaker,” coordinating the actions of other proteins in the MMR pathway, allowing the removal of the mismatch and resynthesis of the resulting gap (reviewed in references 42 and 60). MutS and MutL of E. coli and B. subtilis physically interact with the β clamp (45, 46, 51, 64). Interaction between the B. subtilis β clamp and MutS is important for efficient MMR and organization of MutS-green fluorescent protein (GFP) into foci in response to replication errors, while the function of MutL binding to the β clamp is unknown (64).These studies show that the β clamp is critical for several aspects of DNA metabolism in E. coli and B. subtilis. In E. coli, many dnaN alleles have been examined and used to define the mechanistic roles of the β clamp in vivo (5, 18, 24, 30, 31, 48, 49, 73). A limitation in studying the mechanistic roles of the B. subtilis β clamp is that only two dnaN alleles (β clamp) are available, dnaN5 and dnaN34 (36) (www.bgsc.org/), and both of these alleles do not support growth at temperatures above 49°C, suggesting that they may cause similar defects (36) (www.bgsc.org/). Of these two dnaN alleles, only dnaN5 has been investigated in any detail (36, 53, 64). The mutant β clamp encoded by dnaN5 contains a G73R substitution [dnaN5(G73R)] in a surface-exposed residue located on the outside rim of the β clamp (53, 64). Our previous studies with this allele showed that dnaN5(G73R) confers an increase in mutation frequency at 30°C and 37°C (64). Further characterization of dnaN5(G73R) showed that the increased mutation frequency is caused by a partial defect in MMR (64). Additionally, dnaN5(G73R)-containing cells have a reduced ability to support MutS-GFP focus formation in response to mismatches (64). These results support the hypothesis that G73R in the β clamp causes a defect in DNA replication at 49°C (36) and impaired MMR manifested by a defect in establishing the assembly of MutS-GFP foci in response to replication errors (64).To understand the roles of the B. subtilis β clamp in MMR and DNA replication, we examined the dnaN5 and dnaN34 alleles. We found that the nucleotide sequences of dnaN5 and dnaN34 and the phenotypes they produce were identical, both producing the G73R missense mutation. We analyzed in vivo β clampG73R protein levels and found that the β clampG73R protein accumulated to wild-type levels at elevated temperatures. To identify amino acid residues that would restore DNA replication at elevated temperatures, we isolated three intragenic suppressors of dnaN5(G73R) that conferred growth of B. subtilis cells at 49°C. Epistasis analysis and determination of the mutation spectrum showed that each dnaN allele isolated in this study caused an MMR-dependent increase in mutation frequency. Additionally, we found that the β clamp binding protein YabA can reduce the efficiency of MMR in vivo when yabA expression is induced. Thus, we have identified residues in the β clamp that are critical for DNA replication and MMR in B. subtilis. We also found that a β clamp binding protein, YabA, can reduce the efficiency of MMR in vivo.  相似文献   

7.
8.
9.
10.
An N-carbamoyl-β-alanine amidohydrolase of industrial interest from Agrobacterium tumefaciens C58 (βcarAt) has been characterized. βcarAt is most active at 30°C and pH 8.0 with N-carbamoyl-β-alanine as a substrate. The purified enzyme is completely inactivated by the metal-chelating agent 8-hydroxyquinoline-5-sulfonic acid (HQSA), and activity is restored by the addition of divalent metal ions, such as Mn2+, Ni2+, and Co2+. The native enzyme is a homodimer with a molecular mass of 90 kDa from pH 5.5 to 9.0. The enzyme has a broad substrate spectrum and hydrolyzes nonsubstituted N-carbamoyl-α-, -β-, -γ-, and -δ-amino acids, with the greatest catalytic efficiency for N-carbamoyl-β-alanine. βcarAt also recognizes substrate analogues substituted with sulfonic and phosphonic acid groups to produce the β-amino acids taurine and ciliatine, respectively. βcarAt is able to produce monosubstituted β2- and β3-amino acids, showing better catalytic efficiency (kcat/Km) for the production of the former. For both types of monosubstituted substrates, the enzyme hydrolyzes N-carbamoyl-β-amino acids with a short aliphatic side chain better than those with aromatic rings. These properties make βcarAt an outstanding candidate for application in the biotechnology industry.N-Carbamoyl-β-alanine amidohydrolase (NCβAA) (EC 3.5.1.6), also known as β-alanine synthase or β-ureidopropionase, catalyzes the third and final step of reductive pyrimidine degradation. In this reaction, N-carbamoyl-β-alanine or N-carbamoyl-β-aminoisobutyric acid is irreversibly hydrolyzed to CO2, NH3, and β-alanine or β-aminoisobutyric acid, respectively (43). Eukaryotic NCβAAs have been purified from several sources (10, 25, 33, 39, 42, 44). Nevertheless, only two prokaryotic NCβAAs, belonging to the Clostridium and Pseudomonas genera (4, 29), have been purified to date, although this activity has been inferred for several microorganisms due to the appearance of the reductive pathway of pyrimidine degradation (38, 45). Pseudomonas NCβAA is also able to hydrolyze l-N-carbamoyl-α-amino acids, and indeed, this activity is widespread in the bacterial kingdom (3, 23, 26, 46).β-Amino acids have unique pharmacological properties, and their utility as building blocks of β-peptides, pharmaceutical compounds, and natural products is of growing interest (14). β-Alanine, a natural β-amino acid, is a precursor of coenzyme A and pantothenic acid in bacteria and fungi (vitamin B5) (7). β-Alanine is widely distributed in the central nervous systems of vertebrates and is a structural analogue of γ-amino-n-butyric acid and glycine, major inhibitory neurotransmitters, suggesting that it may be involved in synaptic transmissions (20). Another important natural β-amino acid is taurine (2-aminoethanesulfonic acid), which plays an important role in several essential processes, such as membrane stabilization, osmoregulation, glucose metabolism, antioxidation, and development of the central nervous system and the retina (9, 28, 33). 2-Aminoethylphosphonate, the most common naturally occurring phosphonate, also known as ciliatine, is an important precursor used in the biosynthesis of phosphonolipids, phosphonoproteins, and phosphonoglycans (5). β-Homoalanine (β-aminobutyric acid) has been used successfully for the design of nonnatural ligands for therapeutic application against autoimmune diseases such as rheumatoid arthritis, multiple sclerosis, or autoimmune uveitis (30). Substituted β-amino acids can be denominated β2, β3, and β2,3, depending on the position of the side chain(s) (R) on the amino acid skeleton (18). β2-Amino acids are not yet as readily available as their β3-counterparts, as they must be prepared using multistep procedures (17).We decided to characterize NCβAA (β-carbamoylase) from Agrobacterium tumefaciens C58 (βcarAt) after showing that some dihydropyrimidinases belonging to the Arthrobacter and Sinorhizobium genera are able to hydrolyze different 5- or 6-substituted dihydrouracils to the corresponding N-carbamoyl-β-amino acids (18, 22). If βcarAt could decarbamoylate the reaction products of dihydrouracils, different β-amino acids would be obtained enzymatically in the same way that α-amino acids are produced via the hydantoinase process (6, 21). We therefore describe the physical, biochemical, kinetic, and substrate specificity properties of recombinant βcarAt.  相似文献   

11.
An extracellular β-fructofuranosidase from the yeast Xanthophyllomyces dendrorhous was characterized biochemically, molecularly, and phylogenetically. This enzyme is a glycoprotein with an estimated molecular mass of 160 kDa, of which the N-linked carbohydrate accounts for 60% of the total mass. It displays optimum activity at pH 5.0 to 6.5, and its thermophilicity (with maximum activity at 65 to 70°C) and thermostability (with a T50 in the range 66 to 71°C) is higher than that exhibited by most yeast invertases. The enzyme was able to hydrolyze fructosyl-β-(2→1)-linked carbohydrates such as sucrose, 1-kestose, or nystose, although its catalytic efficiency, defined by the kcat/Km ratio, indicates that it hydrolyzes sucrose approximately 4.2 times more efficiently than 1-kestose. Unlike other microbial β-fructofuranosidases, the enzyme from X. dendrorhous produces neokestose as the main transglycosylation product, a potentially novel bifidogenic trisaccharide. Using a 41% (wt/vol) sucrose solution, the maximum fructooligosaccharide concentration reached was 65.9 g liter−1. In addition, we isolated and sequenced the X. dendrorhous β-fructofuranosidase gene (Xd-INV), showing that it encodes a putative mature polypeptide of 595 amino acids and that it shares significant identity with other fungal, yeast, and plant β-fructofuranosidases, all members of family 32 of the glycosyl-hydrolases. We demonstrate that the Xd-INV could functionally complement the suc2 mutation of Saccharomyces cerevisiae and, finally, a structural model of the new enzyme based on the homologous invertase from Arabidopsis thaliana has also been obtained.The basidiomycetous yeast Xanthophyllomyces dendrorhous (formerly Phaffia rhodozyma) produces astaxanthin (3-3′-dihydroxy-β,β-carotene-4,4 dione [17, 25]). Different industries have displayed great interest in this carotenoid pigment due to its attractive red-orange color and antioxidant properties, which has intensified the molecular and genetic study of this yeast. As a result, several genes involved in the astaxanthin biosynthetic pathway have been cloned and/or characterized, as well as some other genes such as those encoding actin (60), glyceraldehyde-3-phosphate dehydrogenase (56), endo-β-1,3-glucanase, and aspartic protease (4). In terms of the use of carbon sources, a β-amylase (9), and an α-glucosidase (33) with glucosyltransferase activity (12), as well as a yeast cell-associated invertase (41), have also been reported.Invertases or β-fructofuranosidases (EC 3.2.1.26) catalyze the release of β-fructose from the nonreducing termini of various β-d-fructofuranoside substrates. Yeast β-fructofuranosidases have been widely studied, including that of Saccharomyces cerevisiae (11, 14, 45, 46), Schizosaccharomyces pombe (36), Pichia anomala (40, 49), Candida utilis (5, 8), or Schwanniomyces occidentalis (2). They generally exhibit strong similarities where sequences are available, and they have been classified within family 32 of the glycosyl-hydrolases (GH) on the basis of their amino acid sequences. The catalytic mechanism proposed for the S. cerevisiae enzyme implies that an aspartate close to the N terminus (Asp-23) acts as a nucleophile, and a glutamate (Glu-204) acts as the acid/base catalyst (46). In addition, the three-dimensional structures of some enzymes in this family have been resolved, such as that of an exoinulinase from Aspergillus niger (var. awamori; 37) and the invertase from Arabidopsis thaliana (55).As well as hydrolyzing sucrose, β-fructofuranosidases from microorganisms may also catalyze the synthesis of short-chain fructooligosaccharides (FOS), in which one to three fructosyl moieties are linked to the sucrose skeleton by different glycosidic bonds depending on the source of the enzyme (3, 52). FOS are one of the most promising ingredients for functional foods since they act as prebiotics (44), and they exert a beneficial effect on human health, participating in the prevention of cardiovascular diseases, colon cancer, or osteoporosis (28). Currently, Aspergillus fructosyltransferase is the main industrial producer of FOS (15, 52), producing a mixture of FOS with an inulin-type structure, containing β-(2→1)-linked fructose-oligomers (1F-FOS: 1-kestose, nystose, or 1F-fructofuranosylnystose). However, there is certain interest in the development of novel molecules that may have better prebiotic and physiological properties. In this context, β-(2→6)-linked FOS, where this link exits between two fructose units (6F-FOS: 6-kestose) or between fructose and the glucosyl moiety (6G-FOS: neokestose, neonystose, and neofructofuranosylnystose), may have enhanced prebiotic properties compared to commercial FOS (29, 34, 54). The enzymatic synthesis of 6-kestose and other related β-(2→6)-linked fructosyl oligomers has already been reported in yeasts such as S. cerevisiae (11) or Schwanniomyces occidentalis (2) and in fungi such as Thermoascus aurantiacus (26) or Sporotrichum thermophile (27). However, the production of FOS included in the 6G-FOS series has not been widely reported in microorganisms, probably because they are not generally produced (2, 15) or because they represent only a minor biosynthetic product (e.g., with baker''s yeast invertase) (11). Most research into neo-FOS production has been carried out with Penicillium citrinum cells (19, 31, 32, 39). In this context, neokestose is the main transglycosylation product accumulated by whole X. dendrorhous cells from sucrose (30), although the enzyme responsible for this reaction remained uncharacterized.Here, we describe the molecular, phylogenetic, and biochemical characterization of an extracellular β-fructofuranosidase from X. dendrorhous. Kinetic studies of its hydrolytic activity were performed using different substrates, and we investigated its fructosyltransferase capacity. The functionality of the gene analyzed was verified through its heterologous expression, and a structural model of this enzyme based on the homologous invertase from A. thaliana has also been obtained.  相似文献   

12.
Phenoxyalkanoic acid (PAA) herbicides are widely used in agriculture. Biotic degradation of such herbicides occurs in soils and is initiated by α-ketoglutarate- and Fe2+-dependent dioxygenases encoded by tfdA-like genes (i.e., tfdA and tfdAα). Novel primers and quantitative kinetic PCR (qPCR) assays were developed to analyze the diversity and abundance of tfdA-like genes in soil. Five primer sets targeting tfdA-like genes were designed and evaluated. Primer sets 3 to 5 specifically amplified tfdA-like genes from soil, and a total of 437 sequences were retrieved. Coverages of gene libraries were 62 to 100%, up to 122 genotypes were detected, and up to 389 genotypes were predicted to occur in the gene libraries as indicated by the richness estimator Chao1. Phylogenetic analysis of in silico-translated tfdA-like genes indicated that soil tfdA-like genes were related to those of group 2 and 3 Bradyrhizobium spp., Sphingomonas spp., and uncultured soil bacteria. Soil-derived tfdA-like genes were assigned to 11 clusters, 4 of which were composed of novel sequences from this study, indicating that soil harbors novel and diverse tfdA-like genes. Correlation analysis of 16S rRNA and tfdA-like gene similarity indicated that any two bacteria with D > 20% of group 2 tfdA-like gene-derived protein sequences belong to different species. Thus, data indicate that the soil analyzed harbors at least 48 novel bacterial species containing group 2 tfdA-like genes. Novel qPCR assays were established to quantify such new tfdA-like genes. Copy numbers of tfdA-like genes were 1.0 × 106 to 65 × 106 per gram (dry weight) soil in four different soils, indicating that hitherto-unknown, diverse tfdA-like genes are abundant in soils.Phenoxyalkanoic acid (PAA) herbicides such as MCPA (4-chloro-2-methyl-phenoxyacetic acid) and 2,4-D (2,4-dichlorophenoxyacetic acid) are widely used to control broad-leaf weeds in agricultural as well as nonagricultural areas (19, 77). Degradation occurs primarily under oxic conditions in soil, and microorganisms play a key role in the degradation of such herbicides in soil (62, 64). Although relatively rapidly degraded in soil (32, 45), both MCPA and 2,4-D are potential groundwater contaminants (10, 56, 70), accentuating the importance of bacterial PAA herbicide-degrading bacteria in soils (e.g., references 3, 5, 6, 20, 41, 59, and 78).Degradation can occur cometabolically or be associated with energy conservation (15, 54). The first step in the degradation of 2,4-D and MCPA is initiated by the product of cadAB or tfdA-like genes (29, 30, 35, 67), which constitutes an α-ketoglutarate (α-KG)- and Fe2+-dependent dioxygenase. TfdA removes the acetate side chain of 2,4-D and MCPA to produce 2,4-dichlorophenol and 4-chloro-2-methylphenol, respectively, and glyoxylate while oxidizing α-ketoglutarate to CO2 and succinate (16, 17).Organisms capable of PAA herbicide degradation are phylogenetically diverse and belong to the Alpha-, Beta-, and Gammproteobacteria and the Bacteroidetes/Chlorobi group (e.g., references 2, 14, 29-34, 39, 60, 68, and 71). These bacteria harbor tfdA-like genes (i.e., tfdA or tfdAα) and are categorized into three groups on an evolutionary and physiological basis (34). The first group consists of beta- and gammaproteobacteria and can be further divided into three distinct classes based on their tfdA genes (30, 46). Class I tfdA genes are closely related to those of Cupriavidus necator JMP134 (formerly Ralstonia eutropha). Class II tfdA genes consist of those of Burkholderia sp. strain RASC and a few strains that are 76% identical to class I tfdA genes. Class III tfdA genes are 77% identical to class I and 80% identical to class II tfdA genes and linked to MCPA degradation in soil (3). The second group consists of alphaproteobacteria, which are closely related to Bradyrhizobium spp. with tfdAα genes having 60% identity to tfdA of group 1 (18, 29, 34). The third group also harbors the tfdAα genes and consists of Sphingomonas spp. within the alphaproteobacteria (30).Diverse PAA herbicide degraders of all three groups were identified in soil by cultivation-dependent studies (32, 34, 41, 78). Besides CadAB, TfdA and certain TfdAα proteins catalyze the conversion of PAA herbicides (29, 30, 35). All groups of tfdA-like genes are potentially linked to the degradation of PAA herbicides, although alternative primary functions of group 2 and 3 TfdAs have been proposed (30, 35). However, recent cultivation-independent studies focused on 16S rRNA genes or solely on group 1 tfdA sequences in soil (e.g., references 3-5, 13, and 41). Whether group 2 and 3 tfdA-like genes are also quantitatively linked to the degradation of PAA herbicides in soils is unknown. Thus, tools to target a broad range of tfdA-like genes are needed to resolve such an issue. Primers used to assess the diversity of tfdA-like sequences used in previous studies were based on the alignment of approximately 50% or less of available sequences to date (3, 20, 29, 32, 39, 47, 58, 73). Primers specifically targeting all major groups of tfdA-like genes to assess and quantify a broad diversity of potential PAA degraders in soil are unavailable. Thus, the objectives of this study were (i) to develop primers specific for all three groups of tfdA-like genes, (ii) to establish quantitative kinetic PCR (qPCR) assays based on such primers for different soil samples, and (iii) to assess the diversity and abundance of tfdA-like genes in soil.  相似文献   

13.
14.
In model organisms, resistance to inhibitors of cholinesterase 8 (Ric-8), a G protein α (Gα) subunit guanine nucleotide exchange factor (GEF), functions to orient mitotic spindles during asymmetric cell divisions; however, whether Ric-8A has any role in mammalian cell division is unknown. We show here that Ric-8A and Gαi function to orient the metaphase mitotic spindle of mammalian adherent cells. During mitosis, Ric-8A localized at the cell cortex, spindle poles, centromeres, central spindle, and midbody. Pertussis toxin proved to be a useful tool in these studies since it blocked the binding of Ric-8A to Gαi, thus preventing its GEF activity for Gαi. Linking Ric-8A signaling to mammalian cell division, treatment of cells with pertussis toxin, reduction of Ric-8A expression, or decreased Gαi expression similarly affected metaphase cells. Each treatment impaired the localization of LGN (GSPM2), NuMA (microtubule binding nuclear mitotic apparatus protein), and dynein at the metaphase cell cortex and disturbed integrin-dependent mitotic spindle orientation. Live cell imaging of HeLa cells expressing green fluorescent protein-tubulin also revealed that reduced Ric-8A expression prolonged mitosis, caused occasional mitotic arrest, and decreased mitotic spindle movements. These data indicate that Ric-8A signaling leads to assembly of a cortical signaling complex that functions to orient the mitotic spindle.The cortical capture of astral microtubules is essential to generate the forces needed for mitotic spindle positioning for both symmetric and asymmetric cell divisions (23, 29). Failure to either capture astral microtubules or the inappropriate application of pulling forces adversely affects mitotic spindle orientation, and can impede embryogenesis and alter cell fate decisions. Studies examining mitotic spindle orientation in Drosophila embryonic and larval neuroblasts have identified two critical pathways, the Gα/Pins/Mud pathway and the Pins/Dlg/Khc73 pathway (29). The heterotrimeric G-protein α subunit (Gα), Pins (Partner-of-Inscuteable), and Mud (Mushroom body defect) are members of an evolutionarily conserved noncanonical G-protein signaling pathway, which form a tripartite protein complex linked to the apical Par complex by the adapter protein Inscuteable (29, 37). Reducing the level of Gαi, Pins, or Mud prevents neuroblast mitotic spindle alignment. A second spindle orientation pathway involves Pins, the tumor suppressor Discs large (Dlg) and the microtubule plus-end-directed kinesin heavy chain 73 (Khc73). Khc73 binds Dlg and coimmunoprecipitates with Pins. Khc73 localized to astral microtubules can induce Pins-Dlg cortical polarity (27).In canonical G-protein signaling pathways, the binding of ligand to a seven-transmembrane receptor triggers a heterotrimeric G-protein α subunit (Gα) to exchange GTP for GDP, resulting in the dissociation of the Gα subunit from its associated Gβγ heterodimer (12, 20). This exposes interactive sites in the Gα and Gβγ subunits, allowing their binding to and activation of downstream effectors. Since Gα subunits possess an intrinsic GTPase activity, GTP hydrolysis leads to the reassembly of heterotrimeric G protein causing signaling to cease. In noncanonical G-protein signaling the seven-transmembrane receptor is replaced by an intracellular guanine nucleotide exchange factor, such as Ric-8 (37). In studies in Drosophila and Caenorhabditis elegans Ric-8 has been shown to positively regulate Gαi activity and is essential for asymmetric cell divisions (1, 2, 5, 8, 11, 36). Although initially characterized as a guanine nucleotide exchange factor (GEF) for isolated Gαsubunits, more recent biochemical studies have shown that Ric-8A (the mammalian equivalent of Ric-8) also acts on a complex of GDP-Gαi, the mammalian Pins homolog LGN, and NuMA (nuclear mitotic apparatus protein; the mammalian equivalent of Mud) catalytically releasing GTP-Gαi and causing liberation of NuMA from LGN (30, 31). Ric-8A can also catalyze guanine nucleotide exchange on Gαi1 bound to the GPR/GoLoco exchange inhibitor AGS3, a paralog of LGN (33). During mitosis the N-terminal portion of LGN binds NuMA and the C-terminal domain binds GDP-Gαi and the trimolecular complex localizes to the cell cortex, where the dynamic release of NuMA from LGN may regulate aster microtubule pulling during cell division (3, 9, 10, 22).In the present study we examined the role of Ric-8A in mitotic spindle orientation in adherent cells and in polarized MDCK cells. In nonpolarized adherent cells cell such as HeLa, integrin mediated cell-substrate adhesion orients the mitotic spindle parallel to the substratum, and thereby both daughter cells remain attached. This requires the actin cytoskeleton, astral microtubules, the microtubule plus end tracking protein EB1, myosin X, cdc42, LIM kinase 1, and phosphatidylinositol(3,4,5)-triphosphate (PIP3) (13, 18, 32, 34, 35). PIP3 may direct dynein/dynactin-dependent pulling forces on the spindle midcortex to orient the mitotic spindle (34). In polarized cells such as Madin-Darby canine kidney (MDCK) cells, the mitotic spindle is constrained by the topology of the cell and cortical cues provided by adherens junctions (24). In contrast to HeLa cells these cues are insensitive to phosphatidylinositol 3-kinase (PI3K) inhibition, which blocks the generation of PIP3 (34). We found that inhibiting either Ric-8A or Gαi expression impairs the orientation of the metaphase mitotic spindle in HeLa cells and pertussis toxin, which blocks Ric-8A triggered nucleotide exchange, disrupts the normal mitotic spindle alignment of both HeLa and MDCK cells. Impairment of Ric-8A expression or function inhibits the localization of Gαi1, LGN, NuMA, and dynein to the metaphase cortex opposite the spindle poles.  相似文献   

15.
Respiratory syncytial virus (RSV) is a common respiratory viral infection in children which is associated with immune dysregulation and subsequent induction and exacerbations of asthma. We recently reported that treatment of primary human epithelial cells (PHBE cells) with transforming growth factor β (TGF-β) enhanced RSV replication. Here, we report that the enhancement of RSV replication is mediated by induction of cell cycle arrest. These data were confirmed by using pharmacologic inhibitors of cell cycle progression, which significantly enhanced RSV replication. Our data also showed that RSV infection alone resulted in cell cycle arrest in A549 and PHBE cells. Interestingly, our data showed that RSV infection induced the expression of TGF-β in epithelial cells. Blocking of TGF-β with anti-TGF-β antibody or use of a specific TGF-β receptor signaling inhibitor resulted in rescue of the RSV-induced cell cycle arrest, suggesting an autocrine mechanism. Collectively, our data demonstrate that RSV regulates the cell cycle through TGF-β in order to enhance its replication. These findings identify a novel pathway for upregulation of virus replication and suggest a plausible mechanism for association of RSV with immune dysregulation and asthma.Respiratory syncytial virus (RSV) is a single-stranded RNA virus and is a common cause of severe respiratory infections in children. RSV predominantly infects lung epithelial cells, inducing bronchiolitis, and in high-risk individuals it can cause lung fibrosis, airway hyperresponsiveness, mucus secretion, and edema. Interestingly, there is substantial evidence to show that RSV infection induces a dysregulation of the immune response (13, 14, 24, 28, 49). However, the molecular underpinnings of this immune dysregulation are not yet completely understood.It has been established that through its interaction with the immune system, RSV is associated with development and exacerbations of asthma, which is a chronic inflammatory respiratory disease (17, 18, 36, 41). In comparison to healthy individuals, those with asthma have an exaggerated inflammatory response during respiratory virus infections. Despite many studies reporting the involvement of RSV with asthma development and exacerbations, the underlining mechanisms are not yet fully delineated.Previously, we reported that transforming growth factor β (TGF-β) treatment enhanced RSV replication (30). TGF-β is a pleiotropic cytokine with diverse effects on T-cell differentiation and immune regulation and potent anti-inflammatory functions (21, 27, 33, 45). In the lung microenvironment TGF-β inhibits cell proliferation, induces mucus secretion, and regulates airway fibrosis and remodeling (2, 5, 6, 20, 23, 34, 39, 46), all of which are hallmarks of chronic asthma. Specifically, it has been reported that TGF-β expression is elevated in bronchoalveolar lavage fluids and lung tissue of asthmatic patients (9, 32, 48).In addition, genetic studies have found an association between asthma phenotype and TGF-β (19, 26, 38, 43). These studies have identified several single-nucleotide polymorphisms (C509T, T869C, and G915C) in the promoter and coding region of TGF-β that contributed to the increase in gene expression and are significantly associated with childhood wheezing, asthma diagnosis, and asthma severity. Despite this correlation between TGF-β and asthma, the interaction between this key cytokine and respiratory viral infection is poorly understood.A well-known function of TGF-β is the regulation of cell cycle progression. Activation of TGF-β-induced signaling pathways promotes cell cycle arrest in both the G0/G1 and G2/M phases of the cell cycle (7, 8, 25, 29, 40, 42, 44). In the current study, our data showed that TGF-β induction of cell cycle arrest was beneficial to RSV replication. The association of cell cycle arrest with RSV replication was determined by using three different pharmacological inhibitors of cell cycle progression, which enhanced RSV replication. Interestingly, RSV infection alone resulted in secretion of active TGF-β. Treatment of epithelial cells with anti-TGF-β or a specific inhibitor of TGF-β receptor (TGF-βR) signaling resulted in a reduction in RSV replication.In the current study, our data uncover a new pathway for virus regulation of the cell cycle. These findings support our hypothesis that RSV regulates and utilizes TGF-β in lung epithelium to enhance its replication, which may contribute to the physiological changes in the lung leading to immune dysregulation, asthma development, and exacerbations.  相似文献   

16.
Fifteen nonrepetitive ampicillin-resistant Salmonella spp. were identified among 91 Salmonella sp. isolates during nationwide surveillance of Salmonella in waste from 131 chicken farms during 2006 and 2007. Additional phenotyping and genetic characterization of these 15 isolates by using indicator cephalosporins demonstrated that resistance to ampicillin and reduced susceptibility to cefoxitin in three isolates was caused by TEM-1 and DHA-1 β-lactamases. Plasmid profiling and Southern blot analysis of these three DHA-1-positive Salmonella serovar Indiana isolates and previously reported unrelated clinical isolates of DHA-1-positive Salmonella serovar Montevideo, Klebsiella pneumoniae, and Escherichia coli from humans and swine indicated the involvement of the large-size plasmid. Restriction enzyme digestion of the plasmids from the transconjugants showed variable restriction patterns except for the two Salmonella serovar Indiana isolates identified in this study. To the best of our knowledge, this is the first report of the presence of the DHA-1 gene among Salmonella spp. of animal origin.Nontyphoidal Salmonella (NTS) strains are a significant cause of gastrointestinal infections of food origin. These microbes are a heterogeneous group of medically important Gram-negative bacteria and can infect a wide range of animals, including humans (3, 6, 9-11, 25).Currently, no antimicrobial therapies are recommended for the treatment of NTS infection unless a patient is of extreme age, has an underlying disease, or is infected with an invasive Salmonella sp. However, the use of antibiotics in treatment of clinical enteric infection has been heavily compromised by emerging multidrug-resistant microbes (4, 17, 18, 23). In particular, resistance due to extended-spectrum β-lactamases (ESBLs) and AmpC β-lactamases is of special concern as these enzymes confer resistance to some of the front-line antibiotics used to treat enteric infection in humans and animals (4, 13, 14, 19).Four classes of β-lactamases are known to confer resistance to β-lactam antibiotics. Among these, plasmid-mediated class A and class C β-lactamases have been frequently reported, whereas class B and class D β-lactamases are relatively rare (4). TEM and SHV enzymes of class A β-lactamases are generally found in Gram-negative bacteria and are derived by one or more amino acid substitutions around the active site of the enzyme that is responsible for the ESBL phenotype (4). Recently, the CTX-M enzyme of class A β-lactamases has been increasingly reported from enteric microbes, like Salmonella and Escherichia coli (4, 5, 9, 15). These have greater activity against cefotaxime than do other oxyimino-β-lactam substrates, like ceftazidime, ceftriaxone, or cefepime (4, 5). Plasmid-mediated AmpC β-lactamases, like DHA and CMY, are not inhibited by clavulanic acid and have been isolated from a wide variety of clinical and community-acquired microbes (2, 4, 13, 14, 16). These β-lactamases are native to the chromosomes of many Gram-negative bacilli but are missing in some genera, like Salmonella (4). The majority of β-lactamases reported in Salmonella to date have been derived from human clinical isolates, and only limited information is available regarding Salmonella spp. derived from farm animals, although isolates from both humans and animals are of clinical and epidemiological importance (4, 15, 25).In light of this knowledge gap, our study focused on assessing the distribution of Salmonella serovars in poultry farms in South Korea. Subsequently, isolates were analyzed for resistance to antibiotics commonly used in farms. Phenotypic and genetic characteristics of ampicillin-resistant Salmonella isolates were tested to gain insight into what β-lactamases were prevalent among these strains. We also characterized DHA-1-associated plasmids in these Salmonella spp. and compared them with clinical isolates of Salmonella, Klebsiella pneumoniae, and Escherichia coli from humans and from swine.  相似文献   

17.
Endothelial cell (EC) migration, cell-cell adhesion, and the formation of branching point structures are considered hallmarks of angiogenesis; however, the underlying mechanisms of these processes are not well understood. Lipid phosphate phosphatase 3 (LPP3) is a recently described p120-catenin-associated integrin ligand localized in adherens junctions (AJs) of ECs. Here, we tested the hypothesis that LPP3 stimulates β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) to induce EC migration and formation of branching point structures. In subconfluent ECs, LPP3 induced expression of fibronectin via β-catenin/LEF-1 signaling in a phosphatase and tensin homologue (PTEN)-dependent manner. In confluent ECs, depletion of p120-catenin restored LPP3-mediated β-catenin/LEF-1 signaling. Depletion of LPP3 resulted in destabilization of β-catenin, which in turn reduced fibronectin synthesis and deposition, which resulted in inhibition of EC migration. Accordingly, reexpression of β-catenin but not p120-catenin in LPP3-depleted ECs restored de novo synthesis of fibronectin, which mediated EC migration and formation of branching point structures. In confluent ECs, however, a fraction of p120-catenin associated and colocalized with LPP3 at the plasma membrane, via the C-terminal cytoplasmic domain, thereby limiting the ability of LPP3 to stimulate β-catenin/LEF-1 signaling. Thus, our study identified a key role for LPP3 in orchestrating PTEN-mediated β-catenin/LEF-1 signaling in EC migration, cell-cell adhesion, and formation of branching point structures.Angiogenesis, the formation of new blood vessels, involves several well-coordinated cellular processes, including endothelial cell (EC) migration, synthesis and deposition of extracellular matrix proteins, such as fibronectin, cell-cell adhesion, and formation of branching point structures (1-3, 19, 33); however, less is known about the underlying mechanisms of these processes (6, 8, 12, 14, 16, 17). For example, adherens junctions (AJs), which mediate cell-cell adhesion between ECs, may be involved in limiting the extent of cell migration (2, 14, 38, 40). VE-cadherin, a protein found in AJs, is a single-pass transmembrane polypeptide responsible for calcium-dependent homophilic interactions through its extracellular domains (2, 38, 40). The VE-cadherin cytoplasmic domain interacts with the Armadillo domain-containing proteins, β-catenin, γ-catenin (plakoglobin), and p120-catenin (p120ctn) (2, 15, 38, 40, 43). Genetic and biochemical evidence documents a crucial role of β-catenin in regulating cell adhesion as well as proliferation secondary to the central position of β-catenin in the Wnt signaling pathway (13, 16, 25, 31, 44). In addition, the juxtamembrane protein p120ctn regulates AJ stability via binding to VE-cadherin (2, 7, 9, 15, 21, 28, 32, 43). The absence of regulation or inappropriate regulation of β-catenin and VE-cadherin functions is linked to cardiovascular disease and tumor progression (2, 6).We previously identified lipid phosphate phosphatase 3 (LPP3), also known as phosphatidic acid phosphatase 2b (PAP2b), in a functional assay of angiogenesis (18, 19, 41, 42). LPP3 not only exhibits lipid phosphatase activity but also functions as a cell-associated integrin ligand (18, 19, 35, 41, 42). The known LPPs (LPP1, LPP2, and LPP3) (20-23) are six transmembrane domain-containing plasma membrane-bound enzymes that dephosphorylate sphingosine-1-phosphate (S1P) and its structural homologues, and thus, these phosphatases generate lipid mediators (4, 5, 23, 35, 39). All LPPs, which contain a single N-glycosylation site and a putative lipid phosphatase motif, are situated such that their N and C termini are within the cell (4, 5, 22, 23, 35, 39). Only the LPP3 isoform contains an Arg-Gly-Asp (RGD) sequence in the second extracellular loop, and this RGD sequence enables LPP3 to bind integrins (18, 19, 22). Transfection experiments with green fluorescent protein (GFP)-tagged LPP1 and LPP3 showed that LPP1 is apically sorted, whereas LPP3 colocalized with E-cadherin at cell-cell contact sites with other Madin-Darby canine kidney (MDCK) cells (22). Mutagenesis and domain swapping experiments established that LPP1 contains an apical targeting signal sequence (FDKTRL) in its N-terminal segment. In contrast, LPP3 contains a dityrosine (109Y/110Y) basolateral sorting motif (22). Interestingly, conventional deletion of Lpp3 is embryonic lethal, since the Lpp3 gene plays a critical role in extraembryonic vasculogenesis independent of its lipid phosphatase activity (11). In addition, an LPP3-neutralizing antibody was shown to prevent cell-cell interactions (19, 42) and angiogenesis (42). Here, we addressed the hypothesis that LPP3 plays a key role in EC migration, cell-cell adhesion, and formation of branching point structures by stimulating β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) signaling.  相似文献   

18.
The γ134.5 protein of herpes simplex virus 1 is an essential factor for viral virulence. In infected cells, this viral protein prevents the translation arrest mediated by double-stranded RNA-dependent protein kinase R. Additionally, it associates with and inhibits TANK-binding kinase 1, an essential component of Toll-like receptor-dependent and -independent pathways that activate interferon regulatory factor 3 and cytokine expression. Here, we show that γ134.5 is required to block the maturation of conventional dendritic cells (DCs) that initiate adaptive immune responses. Unlike wild-type virus, the γ134.5 null mutant stimulates the expression of CD86, major histocompatibility complex class II (MHC-II), and cytokines such as alpha/beta interferon in immature DCs. Viral replication in DCs inversely correlates with interferon production. These phenotypes are also mirrored in a mouse ocular infection model. Further, DCs infected with the γ134.5 null mutant effectively activate naïve T cells whereas DCs infected with wild-type virus fail to do so. Type I interferon-neutralizing antibodies partially reverse virus-induced upregulation of CD86 and MHC-II, suggesting that γ134.5 acts through interferon-dependent and -independent mechanisms. These data indicate that γ134.5 is involved in the impairment of innate immunity by inhibiting both type I interferon production and DC maturation, leading to defective T-cell activation.Herpes simplex virus 1 (HSV-1) is a human pathogen responsible for localized mucocutaneous lesions and encephalitis (51). Following primary infection, HSV-1 establishes a latent or lytic infection in which the virus interacts with host cells, which include dendritic cells (DCs), required to initiate adaptive immune responses (36). Immature DCs, which reside in almost all peripheral tissues, are able to capture and process viral antigens (15). In this process, DCs migrate to lymph nodes, where they mature and present antigens to T cells. Mature DCs display high levels of major histocompatibility complex class II (MHC-II) and costimulatory molecules such as CD40, CD80, and CD86. Additionally, DCs release proinflammatory cytokines such as interleukin-12 (IL-12), tumor necrosis factor alpha, alpha interferon (IFN-α), and IFN-β, which promote DC maturation and activation. An important feature of functional DCs is to activate naïve T cells, and myeloid submucosal and lymph node resident DCs are responsible for HSV-specific T-cell activation (2, 45, 52). Moreover, DCs play a direct role in innate antiviral immunity by secreting type I IFN.HSV-1 is capable of infecting both immature and mature DCs (20, 24, 34, 38, 42). A previous study suggested that HSV entry into DCs requires viral receptors HVEM and nectin-2 (42). Upon HSV infection, plasmacytoid DCs detect viral genome through Toll-like receptor 9 (TLR9) and produce high levels of IFN-α (16, 23, 30, 40). In contrast, myeloid DCs, which are major antigen-presenting cells, recognize viral components through distinct pathways, independently of TLR9 (16, 36, 40). It has been suggested previously that HSV proteins or RNA intermediates produced during viral replication trigger myeloid DCs (16, 40). Indeed, a protein complex that consists of HSV glycoproteins B, D, H, and L stimulates the expression of CD40, CD83, CD86, and cytokines in myeloid DCs (41). Hence, DCs sense HSV through TLR-dependent and -independent mechanisms (16, 40, 41). Nevertheless, HSV replication compromises DC functions and subsequent T-cell activation (3, 20, 24, 42). HSV-1 interaction with immature DCs results in the downregulation of costimulatory molecules and cytokines (20, 34, 38, 42). While HSV-2 induces rapid apoptosis, HSV-1 does so with a delayed kinetics in human DCs (4, 20, 38). HSV-1 is also reported to interfere with functions of mature DCs (24, 39). Upon infection, HSV-1 induces the degradation of CD83 but not CD80 or CD86 in mature DCs (24, 25). Additionally, HSV-1 reduces levels of the chemokine receptors CCR7 and CXCR4 on mature DCs and subsequently impairs DC migration to the respective chemokine ligands CCL19 and CXCL12 (39).Although HSV infection impairs DC functions, viral components responsible for this impairment have not been thoroughly investigated. It has been suggested previously that the virion host shut-off protein (vhs) of HSV-1 contributes partially to the viral block of DC activation (43). This activity is thought to stem from the ability of vhs to destabilize host mRNA. Emerging evidence suggests that ICP0 perturbs the function of mature DCs, where it mediates CD83 degradation via cellular proteasomes (25). Findings from related studies show that ICP0 inhibits the induction of IFN-stimulated genes mediated by IFN regulatory factor 3 (IRF3) or IRF7 in other cell types (11, 27, 32, 33). However, the link of ICP0 activities to DC maturation remains to be established. Recently, we found that γ134.5, an HSV virulence factor, associates with and inhibits TANK-binding kinase 1 (TBK1), an essential component of TLR-dependent and -independent pathways that activates IRF3 and cytokine expression (49). Interestingly, an HSV mutant lacking γ134.5 stimulates MHC-II surface expression in glioblastoma cells (47). These observations raise the hypothesis that γ134.5 may modulate DC maturation during HSV infection.In this study, we report that γ134.5 is required to perturb DC maturation during HSV infection, leading to impaired T-cell activation. Wild-type virus, but not the γ134.5 null mutant, suppresses the expression of costimulatory molecules as well as cytokines in DCs. We provide evidence that the viral block of DC maturation is associated with reduced IFN-α/β secretion. Furthermore, the expression of γ134.5 inhibits DC-mediated allogeneic T-cell activation and IFN-γ production. IFN-neutralizing antibodies partially reverse DC maturation induced by the γ134.5 null mutant. These results shed light on the role of γ134.5 relevant to DC maturation and T-cell responses in HSV infection.  相似文献   

19.
Protein kinase B (PKB)/Akt is considered to be a key target downstream of insulin receptor substrate 2 (IRS2) in the regulation of β-cell mass. However, while deficiency of IRS2 in mice results in diabetes with insulin resistance and severe failure of β-cell mass and function, only loss of the PKBβ isoform leads to a mild metabolic phenotype with insulin resistance. Other isoforms were reported not to be required for metabolic regulation. To clarify the roles of the three PKB isoforms in the regulation of islet mass and glucose homeostasis, we assessed the metabolic and pancreatic phenotypes of Pkbα, Pkbβ, and Pkbγ-deficient mice. Our study uncovered a novel role for PKBα in the regulation of glucose homeostasis, whereas it confirmed that Pkbβ−/ mice are insulin resistant with compensatory increase of islet mass. Pkbα−/ mice displayed an opposite phenotype with improved insulin sensitivity, lower blood glucose, and higher serum glucagon concentrations. Pkbγ−/ mice did not show metabolic abnormalities. Additionally, our signaling analyses revealed that PKBα, but not PKBβ or PKBγ, is specifically activated by overexpression of IRS2 in β-cells and is required for IRS2 action in the islets.Adaptation of pancreatic islet mass and function relative to metabolic demand maintains glucose homeostasis and may prevent the development of type 2 diabetes. β-Cell proliferation, apoptosis, growth, and function are tightly regulated by various extracellular factors and intracellular signaling pathways (23, 24, 34). In β-cells, insulin receptor substrate 2 (IRS2) controls maintenance and expansion of islet mass (29, 31, 42). In fact, IRS2-deficient mice are insulin resistant, show β-cell failure and hyperglycemia, and finally develop diabetes (26, 42). In contrast, deficiency of IRS1 only causes insulin resistance without the development of diabetes due to a compensatory increase in functional β-cell mass (1, 38). These observations indicated that IRS2, but not IRS1, is necessary for maintenance and compensatory increase of β-cell mass. Furthermore, experiments with isolated islets revealed that overexpression of IRS2, but not of IRS1, can increase β-cell proliferation and protect cells against high-glucose-induced apoptosis (29). Downstream of IRS2, phosphoinositide 3-kinase (PI3K)-protein kinase B (PKB) signaling is considered to be the critical pathway for the regulation of β-cell mass and function (12, 15, 16, 27). The serine-threonine kinase PKB, also known as Akt, is required for various cellular processes, from the regulation of cell cycle, survival, and growth to glucose and protein metabolism. In mammals, three PKB/Akt isoforms have been characterized and named PKBα/Akt1, PKBβ/Akt2, and PKBγ/Akt3. Although encoded by different genes on different chromosomes, the three isoforms display high homology at the protein level with 80 to 85% identical residues and the same structural organization (43). However, they differ in terms of tissue-specific expression. PKBα is expressed in most tissues and PKBβ is highly expressed in insulin-responsive tissues, whereas PKBγ expression is prominent in the brain and testes (17). All three isoforms are expressed in β-cells (30, 37). The roles of PKB in different tissues have been studied in transgenic-mouse models. While Pkbα−/ and Pkbγ−/ mice show impaired fetal growth and brain development, respectively, glucose homeostasis is unaffected in both models (9, 11, 14, 39, 46). In contrast, Pkbβ−/ mice are insulin resistant and mildly glucose intolerant and have less adipose tissue. Depending on the strain and gender, these mice show either late loss of β-cells followed by the development of diabetes and mild growth deficiency or compensatory increase of β-cell mass without age-dependent progression into overt hyperglycemia (10, 17). These studies suggested that PKBβ is the only isoform playing a role in the regulation of energy homeostasis. On the other hand, constitutive activation of PKBα in β-cells is sufficient to increase growth and proliferation (5, 40), and in INS1 cells it prevents free fatty acid (FFA)-induced apoptosis (44). Furthermore, antagonizing total PKB activity in β-cells by ectopic expression of a kinase-dead mutant causes defects in insulin secretion (4), suggesting that in islets PKB is required mainly for normal function of the β-cells. Although these data support the notion that PKB must play a role in pancreatic β-cells, they are not in line with the stronger metabolic phenotype displayed by IRS2-deficient mice. In fact, PKBα and PKBγ appear not to be required to regulate glucose homeostasis (9, 11, 39), and in the case of Pkbβ−/ mice, even though glucose homeostasis is impaired due to strong peripheral insulin resistance, the overall metabolic phenotype is far less severe than in Irs2−/ mice (10), indicating that the capacity for β-cell compensation is retained in the absence of PKBβ.The aim of this study was to clarify the role of PKB in the regulation of islet mass and to define the relevance of PKB isoforms for IRS2 action in β-cells. Although it had been shown that PKBα is dispensable for the regulation of glucose homeostasis (9, 11), we found lower blood glucose concentrations in Pkbα−/ mice. Based on this observation, we assessed in more detail the metabolic and the endocrine pancreatic phenotypes of Pkbα-, Pkbβ-, or Pkbγ-deficient mice. In addition, glucose uptake into fat cells, insulin secretion, and islet cell proliferation were investigated. Contrary to previous assumptions implying that PKBβ is the only (or at least the main) isoform playing a role in the regulation of glucose metabolism, we present evidence that both PKBα and PKBβ isoforms are required in the periphery for regulation of glucose homeostasis. While we confirmed that Pkbβ−/ mice are insulin resistant and glucose intolerant with compensatory increase of β-cell mass, Pkbα−/ mice showed lower blood glucose levels, were more insulin sensitive, and revealed higher serum glucagon concentrations accompanied by a mild increase in α-cell mass and proliferation. Moreover, our in vitro experiments showed that PKBα is specifically activated by IRS2 in β-cells and that its activation is required for IRS2-induced proliferation in islets.  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