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.
2.
We have optimized the display of the B domain of staphylococcal protein A on the surface of Lactococcus lactis. The maximum binding capacity was estimated at 0.146 μg of antibody per 108 cells and was sustained at 86% after treatment with simulated gastric juice. A tumor necrosis factor alpha (TNF-α)-binding affibody was also displayed and bound TNF-α, which could be useful in the treatment of inflammatory bowel disease.Lactic acid bacteria (LAB) have received considerable attention in recent years, based on their record of safe usage as a constituent of fermented foods and their health-promoting effects as probiotics (5). Recombinant LAB could also be used in therapy, with most applications aimed at the delivery of antigens or therapeutic proteins to human mucosal surfaces (27). Another potential application of recombinant LAB involves surface attachment of affinity molecules, such as antibodies, single-chain variable fragments (scFv), or specific oligosaccharides, which can target pathogens, toxins, or inflammatory mediators in the human intestine (9, 22).We describe the surface display of two types of affinity molecule, the B domain and the tumor necrosis factor alpha (TNF-α)-binding affibody, on a model LAB, Lactococcus lactis. The B domain, which is one out of five antibody-binding domains of staphylococcal protein A (15), was already used as a model protein (2). The surface display of the B domain was reported for Escherichia coli in a biosensor application (8), for Saccharomyces cerevisiae as a whole-cell immunoadsorbent (16), and for certain viruses for specific cell targeting (20).The affinity of the B domain or its synthetic homologue, the Z domain (17), for the antibody Fc region has been redirected to several other proteins by randomization of amino acids involved in the interaction, using the genetic combinatorial library and phage display (18). The variants of the Z domain were termed “affibodies” and were directed against various proteins (reviewed in reference 19) and also against TNF-α (7).TNF-α is well established as a proinflammatory cytokine in the pathology of inflammatory bowel disease (IBD), and monoclonal antibodies against TNF-α are routinely used in parenteral therapy (23) but can have systemic side effects. The abundant presence of TNF-α in the stool samples of IBD patients (4) and the successful treatment of experimental colitis in rats by oral administration of avian IgY (28) justify the oral delivery of an agent with the capability of removing TNF-α in IBD. We have therefore replaced the B domain in our surface-displayed fusion protein with an affibody against TNF-α (7) as a second type of binding molecule. LAB with surface-displayed affibody against TNF-α could be used to bind TNF-α in the intestine, with the potential for use in the treatment of IBD. This novel approach could also overcome the problems with the gastrointestinal stability of antibodies.  相似文献   

3.
4.
5.
6.
7.
Receptor protein tyrosine phosphatase α (RPTPα) is the mitotic activator of the protein tyrosine kinase Src. RPTPα serine hyperphosphorylation was proposed to mediate mitotic activation of Src. We raised phosphospecific antibodies to the two main serine phosphorylation sites, and we discovered that RPTPα Ser204 was almost completely dephosphorylated in mitotic NIH 3T3 and HeLa cells, whereas Ser180 and Tyr789 phosphorylation were only marginally reduced in mitosis. Concomitantly, Src pTyr527 and pTyr416 were dephosphorylated, resulting in 2.3-fold activation of Src in mitosis. Using inhibitors and knockdown experiments, we demonstrated that dephosphorylation of RPTPα pSer204 in mitosis was mediated by PP2A. Mutation of Ser204 to Ala did not activate RPTPα, and intrinsic catalytic activity of RPTPα was not affected in mitosis. Interestingly, binding of endogenous Src to RPTPα was induced in mitosis. GRB2 binding to RPTPα, which was proposed to compete with Src binding to RPTPα, was only modestly reduced in mitosis, which could not account for enhanced Src binding. Moreover, we demonstrate that Src bound to mutant RPTPα-Y789F, lacking the GRB2 binding site, and mutant Src with an impaired Src homology 2 (SH2) domain bound to RPTPα, illustrating that Src binding to RPTPα is not mediated by a pTyr-SH2 interaction. Mutation of RPTPα Ser204 to Asp, mimicking phosphorylation, reduced coimmunoprecipitation with Src, suggesting that phosphorylation of Ser204 prohibits binding to Src. Based on our results, we propose a new model for mitotic activation of Src in which PP2A-mediated dephosphorylation of RPTPα pSer204 facilitates Src binding, leading to RPTPα-mediated dephosphorylation of Src pTyr527 and pTyr416 and hence modest activation of Src.Protein tyrosine phosphatases (PTPs) are responsible for dephosphorylation of the phosphotyrosyl residues. The human genome contains approximately 100 genes that encode members of the four PTP families, and most of them have mouse orthologues (2, 48). According to their subcellular localization, the classical PTPs, encoded by less than half of the total PTP genes, are divided into two subfamilies: cytoplasmic and receptor protein tyrosine phosphatases (RPTPs). The majority of the RPTPs contain, besides a variable extracellular domain and a transmembrane domain, two highly homologous phosphatase domains (27), with the membrane-proximal domain comprising most of the catalytic activity (33).RPTPα is a typical RPTP with a small, highly glycosylated extracellular domain (13). RPTPα function is regulated by many mechanisms, including proteolysis (18), oxidation (55), dimerization (7, 23, 24, 47, 52), and phosphorylation of serine and tyrosine residues (16, 17, 49). RPTPα is broadly expressed in many cell types, and over the years, RPTPα has been shown to be involved in a number of signaling mechanisms, including neuronal (15) and skeletal muscle (34) cell differentiation, neurite elongation (8, 9, 56), insulin receptor signaling downregulation (3, 28, 30, 31, 35), insulin secretion (25), activation of voltage-gated potassium channel Kv1.2 (51), long-term potentiation in hippocampal neurons (32, 38), matrix-dependent force transduction (53), and cell spreading and migration (21, 45, 57).The majority of the roles played in these cellular processes involve RPTPα''s ability to activate the proto-oncogenes Src and Fyn by dephosphorylating their C-terminal inhibitory phosphotyrosine (5, 15, 39, 45, 61). Normally, this phosphotyrosine (pTyr527 in chicken Src) binds to the Src homology 2 (SH2) domain, keeping the protein in an inactive closed conformation. A displacement mechanism was proposed for RPTPα-mediated Src activation in which pTyr789 of RPTPα is required to bind the SH2 domain of Src before RPTPα dephosphorylates Tyr527 (58). This model is the subject of debate since other studies show that RPTPα lacking Tyr789 is still able to dephosphorylate and activate Src (12, 26, 29, 56). In normal cells, Src reaches its activation peak during mitosis (4, 11, 40, 42), and with the help of overexpressing cells, it was shown that this activation is triggered mainly by RPTPα. The model that emerged is that RPTPα is activated in mitosis due to serine hyperphosphorylation and detaches from the GRB2 scaffolding protein (59, 60) that normally binds most of the pTyr789 of RPTPα via its SH2 domain (14, 17, 46). Two serine phosphorylation sites were mapped in the juxtamembrane domain of RPTPα, Ser180 and Ser204 (49). The kinases that were found responsible for their phosphorylation were protein kinase C delta (PKCdelta) (10) and CaMKIIalpha (9), but there is no clear evidence that these kinases are activated in mitosis. We set out to investigate the role of serine phosphorylation of RPTPα in mitotic activation of Src.We generated phosphospecific antibodies and show that RPTPα pSer204, but not pSer180, is dephosphorylated in mitotic NIH 3T3 and HeLa cells, concomitantly with activation of Src. Selective inhibitors suggested that PP2A was the phosphatase that dephosphorylated pSer204. RNA interference (RNAi)-mediated knockdown of the catalytic subunit of PP2A demonstrated that indeed PP2A was responsible for mitotic dephosphorylation of RPTPα pSer204. It is noteworthy that PP2A is known to be activated in mitosis. Intrinsic PTP activities of RPTPα were similar in unsynchronized and mitotic cells, and mutation of Ser204 did not activate RPTPα in in vitro PTP assays. Yet, Src binding to RPTPα was induced in mitotic NIH 3T3 cells and RPTPα-S204D with a phosphomimicking mutation at Ser204 coimmunoprecipitated less efficiently with Src. Based on our results, we propose a mechanism for mitotic activation of Src that is triggered by dephosphorylation of RPTPα pSer204, resulting in enhanced affinity for Src and subsequent dephosphorylation and activation of Src.  相似文献   

8.
9.
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   

10.
Receptors (FcγRs) for the constant region of immunoglobulin G (IgG) are an important link between humoral immunity and cellular immunity. To help define the role of FcγRs in determining the fate of human immunodeficiency virus type 1 (HIV-1) immune complexes, cDNAs for the four major human Fcγ receptors (FcγRI, FcγRIIa, FcγRIIb, and FcγRIIIa) were stably expressed by lentiviral transduction in a cell line (TZM-bl) commonly used for standardized assessments of HIV-1 neutralization. Individual cell lines, each expressing a different FcγR, bound human IgG, as evidence that the physical properties of the receptors were preserved. In assays with a HIV-1 multisubtype panel, the neutralizing activities of two monoclonal antibodies (2F5 and 4E10) that target the membrane-proximal external region (MPER) of gp41 were potentiated by FcγRI and, to a lesser extent, by FcγRIIb. Moreover, the neutralizing activity of an HIV-1-positive plasma sample known to contain gp41 MPER-specific antibodies was potentiated by FcγRI. The neutralizing activities of monoclonal antibodies b12 and 2G12 and other HIV-1-positive plasma samples were rarely affected by any of the four FcγRs. Effects with gp41 MPER-specific antibodies were moderately stronger for IgG1 than for IgG3 and were ineffective for Fab. We conclude that FcγRI and FcγRIIb facilitate antibody-mediated neutralization of HIV-1 by a mechanism that is dependent on the Fc region, IgG subclass, and epitope specificity of antibody. The FcγR effects seen here suggests that the MPER of gp41 could have greater value for vaccines than previously recognized.Fc receptors (FcRs) are differentially expressed on a variety of cells of hematopoietic lineage, where they bind the constant region of antibody (Ab) and provide a link between humoral and cellular immunity. Humans possess two classes of FcRs for the constant region of IgG (FcγRs) that, when cross-linked, are distinguished by their ability to either activate or inhibit cell signaling (69, 77, 79). The activating receptors FcγRI (CD64), FcγRIIa (CD32), and FcγRIII (CD16) signal through an immunoreceptor tyrosine-based activation motif (ITAM), whereas FcγRIIb (CD32) contains an inhibitory motif (ITIM) that counters ITAM signals and B-cell receptor signals. It has been suggested that a balance between activating and inhibitory FcγRs coexpressed on the same cells plays an important role in regulating adaptive immunity (23, 68). Moreover, the inhibitory FcγRIIb, being the sole FcγR on B cells, appears to play an important role in regulating self-tolerance (23, 68). The biologic role of FcγRs may be further influenced by differences in their affinity for immunoglobulin G (IgG); thus, FcγRI is a high-affinity receptor that binds monomeric IgG (mIgG) and IgG immune complexes (IC), whereas FcγRIIa, FcγRIIb, and FcγRIIIa are medium- to low-affinity receptors that preferentially bind IgG IC (10, 49, 78). FcγRs also exhibit differences in their relative affinity for the four IgG subclasses (10), which has been suggested to influence the balance between activating and inhibitory FcγRs (67).In addition to their participation in acquired immunity, FcγRs can mediate several innate immune functions, including phagocytosis of opsonized pathogens, Ab-dependent cell cytotoxicity (ADCC), antigen uptake by professional antigen-presenting cells, and the production of inflammatory cytokines and chemokines (26, 35, 41, 48, 69). In some cases, interaction of Ab-coated viruses with FcγRs may be exploited by viruses as a means to facilitate entry into FcγR-expressing cells (2, 33, 47, 84). Several groups have reported FcγR-mediated Ab-dependent enhancement (ADE) of HIV-1 infection in vitro (47, 51, 58, 63, 94, 96), whereas other reports have implicated FcγRs in efficient inhibition of the virus in vitro (19, 21, 29, 44-46, 62, 98) and possibly as having beneficial effects against HIV-1 in vivo (5, 27, 28, 42). These conflicting results are further complicated by the fact that HIV-1-susceptible cells, such as monocytes and macrophages, can coexpress more than one FcγR (66, 77, 79).HIV-1 entry requires sequential interactions between the viral surface glycoprotein, gp120, and its cellular receptor (CD4) and coreceptor (usually CCR5 or CXCR4), followed by membrane fusion that is mediated by the viral transmembrane glycoprotein gp41 (17, 106). Abs neutralize the virus by binding either gp120 or gp41 and blocking entry into cells. Several human monoclonal Abs that neutralize a broad spectrum of HIV-1 variants have attracted considerable interest for vaccine design. Epitopes for these monoclonal Abs include the receptor binding domain of gp120 in the case of b12 (71, 86), a glycan-specific epitope on gp120 in the case of 2G12 (13, 85, 86), and two adjacent epitopes in the membrane-proximal external region (MPER) of g41 in the cases of 2F5 and 4E10 (3, 11, 38, 93). At least three of these monoclonal Abs have been shown to interact with FcRs and to mediate ADCC (42, 43).A highly standardized and validated assay for neutralizing Abs against HIV-1 that quantifies reductions in luciferase (Luc) reporter gene expression after a single round of virus infection in TZM-bl cells has been developed (60, 104). TZM-bl (also called JC53BL-13) is a CXCR4-positive HeLa cell line that was engineered to express CD4 and CCR5 and to contain integrated reporter genes for firefly Luc and Escherichia coli β-galactosidase under the control of the HIV-1 Tat-regulated promoter in the long terminal repeat terminal repeat sequence (74, 103). TZM-bl cells are permissive to infection by a wide variety of HIV-1, simian immunodeficiency virus, and human-simian immunodeficiency virus strains, including molecularly cloned Env-pseudotyped viruses. Here we report the creation and characterization of four new TZM-bl cell lines, each expressing one of the major human FcγRs. These new cell lines were used to gain a better understanding of the individual roles that FcγRs play in determining the fate of HIV-1 IC. Two FcγRs that potentiated the neutralizing activity of gp41 MPER-specific Abs were identified.  相似文献   

11.
12.
13.
14.
15.
16.
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   

17.
18.
19.
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   

20.
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   

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