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 968 毫秒
1.
2.
Eukaryotic DNA polymerase η (Polη) confers ultraviolet (UV) resistance by catalyzing translesion synthesis (TLS) past UV photoproducts. Polη has been studied extensively in budding yeast and mammalian cells, where its interaction with monoubiquitylated proliferating cell nuclear antigen (PCNA) is necessary for its biological activity. Recently, in collaboration with other investigators, our laboratory demonstrated that Arabidopsis thaliana Polη is required for UV resistance in plants. Furthermore, the purified enzyme can perform TLS opposite a cyclobutane pyrimidine dimer and interacts with PCNA. Intriguingly, the biological activity of Polη in a heterologous yeast assay depends on co-expression with Arabidopsis PCNA2 and Polη sequences implicated in binding PCNA or ubiquitin. We suggest that interaction of Arabidopsis Polη with ubiquitylated PCNA2 is required for TLS past UV photoproducts by Polη.Key words: polymerase η, proliferating cell nuclear antigen, translesion synthesis, ubiquitin, Arabidopsis thaliana, ultraviolet radiationUltraviolet (UV)-induced pyrimidine dimers can block the progression of DNA replication forks potentially disrupting the replication machinery and resulting in cell death. For this reason, cells have evolved non-essential, low fidelity DNA polymerases (Pols) capable of copying damaged templates,1,2 a process termed translesion DNA synthesis (TLS). In budding yeast, TLS past UV photoproducts is catalyzed by Polη and Polζ (composed of the Rev3 catalytic and Rev7 accessory subunits), but also involves the Rev1 protein in an as yet undetermined role linked to Polζ.1,3,4 Yeast and human Polη replicates cyclobutane pyrimidine dimers (CPDs), in particular thymine-thymine (TT) CPDs, in a relatively error-free manner whereas Polζ is essential for UV mutagenesis implicating it in error-prone TLS.1,4,5Both UV resistance due to TLS and the polymerases responsible have been well-studied in yeast and mammalian cells over the past decade. Only more recently has evidence emerged that TLS may also contribute to UV resistance in plants. Arabidopsis thaliana POLH, REV1, REV3 and REV7 encode homologs of Polη, Rev1, Rev3 and Rev7, respectively.610 T-DNA insertions in POLH, REV1 or REV3 sensitise root growth to acute UV doses,68,10 and these mutations, as well as inactivation of REV7, increase the sensitivity of whole plants to longer term UV treatment.6,8 Interestingly, polh rev3 double mutants show an additive increase in UV sensitivity over that observed for polh and rev3 single mutants,6,10 potentially pointing to differences in the UV photoproducts bypassed by the two polymerases. That the enhanced UV sensitivity of the mutants may reflect a TLS deficiency is suggested by the finding that purified Arabidopsis Polη catalyzes primer extension and TLS past a TT CPD in vitro.6For TLS to occur, Polη must gain access to the replication machinery arrested at a UV photoproduct. It does so in yeast and mammalian cells by interacting with proliferating cell nuclear antigen (PCNA), the eukaryotic sliding clamp required for processive DNA replication.1,3,11, DNA damage or stalling of the replicative polymerase triggers monoubiquitylation of PCNA at lysine 164 by a complex of the E2 ubiquitin conjugase Rad6 and the E3 ubiquitin ligase Rad18.1,3,11,12 This modification increases the affinity of Polη for PCNA, with which it interacts via a single PCNA interacting peptide (PIP) box and a single ubiquitin-binding zinc finger (UBZ) domain.1,3In contrast to its yeast and mammalian counterparts, Polη from Arabidopsis and Oryza sativa (rice) has two PIP boxes and lacks a UBZ.6,9,10 Instead the two polymerases each possess two ubiquitin-binding motifs (UBMs) similar to those present in the Arabidopsis Rev1 protein and a vertebrate TLS polymerase, Pol., for which there is no homolog in Arabidopsis.6,13 Considerable differences in the sequences flanking the UBMs in Polη and Rev1 argue that Polη did not acquire its UBMs from Rev1, and so, although perhaps unique to plant Polη, their origin remains a mystery.The presence of PCNA- and ubiquitin-binding sequences in plant Polη hint that it may operate in TLS in a manner similar to that for Polη from yeast or mammalian cells. Indeed, three lines of evidence6 lead us to suggest that the Polη PIP boxes and UBMs likely function in binding ubiquitylated PCNA and this interaction is probably required for TLS past UV photoproducts by Arabidopsis Polη. First, Arabidopsis Polη interacts physically and in yeast two-hybrid assays with Arabidopsis PCNA1 and PCNA2. Second, expression in yeast of Arabidopsis cDNAs encoding Polη and PCNA2, but not PCNA1, fully complements the UV sensitivity conferred by elimination of yeast Polη. In vitro mutagenesis suggests the inability of Polη plus PCNA1 to restore UV resistance is due to a lysine at position 201 in PCNA1 but not PCNA2. In the three-dimensional structure of PCNA, amino acid 201 lies adjacent to lysine-164, the residue that is ubiquitylated in yeast and human PCNA. Thus, one possibility is that lysine-201 in PCNA1 prevents complementation of UV sensitivity by inhibiting ubiquitylation of lysine-164. Third, altering presumed critical residues in either of the two PIP boxes or UBM2 in Arabidopsis Polη also prevents restoration of UV resistance in Polη-deficient yeast cells.Several important parts of the puzzle remain to be solved. In particular, the ubiquitylation of plant PCNA has yet to be demonstrated, and the identity of the proteins that might monoubiquitylate plant PCNA is uncertain. Although Arabidopsis Rad6 homologs can ubiquitylate target proteins in vitro, there is no evidence that Arabidopsis PCNA1 or PCNA2 is a substrate, and Arabidopsis lacks a Rad18 homolog.14,15 Finally, if PCNA is ubiquitylated in planta, does this occur at lysine-164 in response to DNA damage or replication fork stalling, is the interaction of Polη with PCNA stimulated by this modification, and is an enhanced interaction mediated by the Polη UBMs?  相似文献   

3.
Reactive oxygen species are ubiquitous mutagens that have been linked to both disease and aging. The most studied oxidative lesion is 7,8-dihydro-8-oxoguanine (GO), which is often miscoded during DNA replication, resulting specifically in GC → TA transversions. In yeast, the mismatch repair (MMR) system repairs GO·A mismatches generated during DNA replication, and the polymerase η (Polη) translesion synthesis DNA polymerase additionally promotes error-free bypass of GO lesions. It has been suggested that Polη limits GO-associated mutagenesis exclusively through its participation in the filling of MMR-generated gaps that contain GO lesions. In the experiments reported here, the SUP4-o forward-mutation assay was used to monitor GC → TA mutation rates in strains defective in MMR (Msh2 or Msh6) and/or in Polη activity. The results clearly demonstrate that Polη can function independently of the MMR system to prevent GO-associated mutations, presumably through preferential insertion of cytosine opposite replication-blocking GO lesions. Furthermore, the Polη-dependent bypass of GO lesions is more efficient on the lagging strand of replication and requires an interaction with proliferating cell nuclear antigen. These studies establish a new paradigm for the prevention of GO-associated mutagenesis in eukaryotes.Eukaryotic genome stability can be compromised by changes at the nucleotide level, alterations in chromosome structure, or changes in chromosome number. Although such changes are responsible for many human diseases, including cancer, a low level of instability is necessary to provide the raw material for evolutionary processes. Changes at the nucleotide level generally occur during replication, either as errors made when copying an undamaged DNA template or during the bypass of DNA lesions. Many types of DNA lesions are due to reactive oxygen species (ROS), which are generated by exposure to physical and chemical mutagens, as well as by normal metabolic processes, such as aerobic respiration (12, 32). Although cells contain multiple antioxidants and other proteins that protect the genome from oxidative damage, ROS have been implicated as causal agents of many diseases and of aging (11, 50).The most common oxidized DNA lesion is 7,8-dihydro-8-oxoguanine, which is referred to here as a GO lesion. The mutagenic potential of this lesion is due to miscoding during DNA synthesis, with replicative DNA polymerases usually misinserting adenine across from the lesion to generate GO·A mispairs and ultimately GC → TA transversions (49). Studies examining the crystal structure of T7 DNA polymerase complexed with a GO·C base pair or a GO·A mispair indicate the basis of this mutagenic specificity. Whereas the GO·C structure physically resembles that of a mismatch, the GO·A mispair structurally resembles a normal Watson-Crick base pair and therefore is likely to escape polymerase-associated proofreading activity (6). A GO-containing nucleotide triphosphate (8-oxo-dGTP) can also be used by DNA polymerases during DNA synthesis, leading specifically to AT > CG transversion events (7).There are three major proteins in Escherichia coli that work independently to prevent GO-associated mutagenesis: MutM (Fpg), MutY, and MutT (36). MutM is a DNA glycosylase that removes GO lesions in the GO·C base pairs created by oxidation of guanine in normal G·C base pairs, while MutY is an adenine-DNA glycosylase that removes adenines from the GO·A mispairs created by incorporation of adenine opposite a GO lesion. If DNA replication occurs before MutM can remove the GO lesion from a GO·C base pair, the lesion will likely generate a GO·A mispair, which is then subjected to the A-specific activity of the MutY protein. Once MutY removes the adenine from the newly synthesized strand, a cytosine can be inserted opposite the lesion, giving MutM another opportunity to excise the GO base. MutT is an 8-oxo-dGTPase that degrades 8-oxo-dGTP, thereby greatly reducing its incorporation into DNA. The postreplicative mismatch repair (MMR) pathway has also been implicated in preventing GO-associated mutagenesis in E. coli by functioning as an alternative to MutY or by helping MutY identify and remove mismatched adenines from GO·A mispairs (3, 60).In the yeast Saccharomyces cerevisiae, the Ogg1 protein is the functional homolog of MutM (55) and thus removes GO lesions that are base paired with cytosine. The MMR machinery is functionally analogous to the MutY protein (37), excising adenines that are inserted opposite GO lesions during DNA replication. The mismatch recognition MutSα complex (a heterodimer of the Msh2 and Msh6 proteins) specifically recognizes GO·A mispairs and initiates removal of the portion of the newly synthesized strand containing the adenine (37). A homolog of MutT has yet to be identified in yeast, although one does exist in mammalian cells (23). It is possible that the MutT homolog has eluded discovery either because it is essential, because there is a redundant activity, or because 8-oxo-dGTP is not a significant mutagen in yeast.A third mechanism that limits GO-associated mutagenesis in yeast involves the translesion synthesis (TLS) polymerase, polymerase η (Polη), which is a member of the Y family of DNA polymerases and is encoded by the RAD30 gene (18, 61). Y family polymerases have a large active-site pocket that can accommodate structurally deformed bases, enabling them to insert a nucleotide opposite a lesion (29). Not only is such lesion bypass potentially error prone, the larger active-site pocket of TLS polymerases imparts very low fidelity when copying undamaged DNA. Polη, for example, is error prone when bypassing some lesions, such as abasic sites (17), but has relatively high fidelity when bypassing GO lesions, usually inserting a cytosine across from the lesion (18, 61). At GO lesions, Polη is 10-fold more accurate and efficient than Polδ (34). When given an undamaged DNA template, however, the base substitution error frequency of Polη in vitro is 3 orders of magnitude greater than that of a typical replicative polymerase (35). In addition to Polη, there are two other TLS polymerases in S. cerevisiae (Polζ and Rev1), but neither has been implicated in the bypass of GO lesions (10, 48).The most straightforward way for Polη to be involved in GO bypass would be for it to be recruited when a replicative polymerase stalls or leaves behind a gap. The replication-blocking potential of GO lesions, however, is unclear. Some in vitro studies have shown that replicative DNA polymerases stall or pause when encountering a template GO lesion (18, 47), but other studies have suggested that this is not the case (49). The currently accepted model is that Polη is specifically recruited to fill the gaps generated by MMR when it initiates correction of GO·A mispairs (18, 54). This model of MMR-Polη cooperation in preventing GO-associated mutagenesis is based on epistasis analysis performed using the CAN1 forward mutation assay (18). Although the relationship between msh2 and rad30 was concluded to be epistatic, the data are also consistent with an additive relationship and, hence, potentially independent roles of Msh2 and Polη in limiting GO-associated mutagenesis. How and why the MMR pathway might specifically recruit a generally error-prone polymerase to fill the gaps in what is normally an extremely accurate repair process is not obvious.In the present study, a SUP4-o forward-mutation system was used to reexamine the relationship between MMR and Polη in preventing GO-associated mutagenesis in yeast. To enhance the accumulation of GO lesions, all experiments were conducted in mutants defective in removing GO from GO·C base pairs (an ogg1 background). In addition, both msh2 and msh6 mutants were analyzed. In an msh6 background, loss of the MutSα heterodimer eliminates the correction of GO·A mispairs, while retention of MutSβ (a heterodimer of Msh2 and Msh3) allows continued correction of most insertion-deletion loops. Finally, mutation spectra, as well as mutation rates, were considered in order to focus specifically on GC → TA mutations. The results reported here demonstrate that Polη can function independently of MMR to prevent GO-associated mutagenesis, presumably through its ability to bypass these lesions in an error-free manner. The data further indicate that the Polη-dependent bypass of GO lesions is more efficient on the lagging strand of replication and that it requires interaction with proliferating cell nuclear antigen (PCNA).  相似文献   

4.
5.
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   

6.
Functional coordination between DNA replication helicases and DNA polymerases at replication forks, achieved through physical linkages, has been demonstrated in prokaryotes but not in eukaryotes. In Saccharomyces cerevisiae, we showed that mutations that compromise the activity of the MCM helicase enhance the physical stability of DNA polymerase α in the absence of their presumed linker, Mcm10. Mcm10 is an essential DNA replication protein implicated in the stable assembly of the replisome by virtue of its interaction with the MCM2-7 helicase and Polα. Dominant mcm2 suppressors of mcm10 mutants restore viability by restoring the stability of Polα without restoring the stability of Mcm10, in a Mec1-dependent manner. In this process, the single-stranded DNA accumulation observed in the mcm10 mutant is suppressed. The activities of key checkpoint regulators known to be important for replication fork stabilization contribute to the efficiency of suppression. These results suggest that Mcm10 plays two important roles as a linker of the MCM helicase and Polα at the elongating replication fork—first, to coordinate the activities of these two molecular motors, and second, to ensure their physical stability and the integrity of the replication fork.The key players of the replication machinery are the DNA polymerases that synthesize the leading and lagging daughter strands and the replicative helicase that unwinds the parental strands ahead of the polymerases. Coordination between the helicase and the polymerases is critical during replication. Uncoupling of these two molecular machines, especially during lagging strand synthesis, may result in an unrestrained helicase and the exposure of extensive single-stranded DNA (ssDNA), as observed in checkpoint mutants treated with hydroxyurea (HU) (37). Although there is no direct evidence, the implication is that the replicative helicase would be moving at a faster pace than would the DNA polymerase if synchrony were destroyed. In Escherichia coli, the replicative helicase (DnaB) and the primase (DnaG) are coupled by direct contact to form a tight complex (3). In T7, processivity of the gp5 polymerase in lagging strand synthesis requires coupling to the gp4 helicase (16). Recent studies of the budding yeast Saccharomyces cerevisiae suggest that Mrc1 may couple DNA polymerase ɛ and the MCM helicase on the leading strand as well as activate the checkpoint response under replication stress (1, 22, 28). A candidate for coupling DNA polymerase α primase and the MCM helicase on the lagging strand is Mcm10, because Mcm10 interacts with subunits of the Mcm2-7 helicase (26, 29) as well as Polα (14, 33) and the stability of Polα requires Mcm10 in both budding yeast and human cells (8, 33). Mcm10 is an essential protein known to be involved in various aspects of the replication process. It is required during both initiation and elongation steps of DNA replication and interacts with a wide range of replication factors, such as ORC (17, 23, 29), MCM helicase, DNA polymerases ɛ and δ (23), Cdc45 (34), and Polα (33). Therefore, Mcm10 is important for the overall stability of the elongation complex, but its essential function remains unknown.Accumulating evidence suggests that the major function of many checkpoint proteins is the stabilization of the replication machinery at the fork (9, 22, 39), in addition to regulation of the temporal and spatial firing of origins and prevention of premature mitosis (31, 35, 39). The main signal that leads to checkpoint activation is believed to be the exposure of RPA-coated ssDNA (42). In Xenopus, ssDNA exposure has been shown to be mediated by a functional uncoupling between the polymerase and the helicase (7), and it has been shown that the level of checkpoint activation depended on the extent of ssDNA accumulation. This observation suggests that uncoupling of the polymerase and the helicase activity would result in ssDNA accumulation that in turn would activate the checkpoint pathway to stabilize the fork.In our study, we carried out a random and a gene-targeted mutagenesis screen to identify mutations that suppress the conditional lethality of mcm10 caused by the lability of Mcm10 in budding yeast (27). We found suppressor mutations in MCM2, which encodes one of the six distinct subunits of the MCM helicase. These mcm2 mutations correct the fork defects of mcm10, particularly that which leads to Polα instability. The altered helicase activity and activation of the checkpoint pathway of the mcm2 mutants appeared to be required for viability of mcm10 mcm2. We showed that uncoupling the MCM helicase and DNA polymerase α by destabilizing Mcm10 leads to accumulation of ssDNA, which is suppressed by reducing the MCM helicase activity. Our findings suggest that the physical coupling of Polα and the helicase by Mcm10 may be replaced by an alternative stabilization mechanism that involves slowing down the helicase and activating the checkpoint proteins.  相似文献   

7.
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   

8.
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   

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.
11.
12.
During infection, simian virus 40 (SV40) attempts to take hold of the cell, while the host responds with various defense systems, including the ataxia-telangiectasia mutated/ATM-Rad3 related (ATM/ATR)-mediated DNA damage response pathways. Here we show that upon viral infection, ATR directly activates the p53 isoform Δp53, leading to upregulation of the Cdk inhibitor p21 and downregulation of cyclin A-Cdk2/1 (AK) activity, which force the host to stay in the replicative S phase. Moreover, downregulation of AK activity is a prerequisite for the generation of hypophosphorylated, origin-competent DNA polymerase α-primase (hypo-Polα), which is, unlike AK-phosphorylated Polα (P-Polα), recruited by SV40 large T antigen (T-Ag) to initiate viral DNA replication. Prevention of the downregulation of AK activity by inactivation of ATR-Δp53-p21 signaling significantly reduced the T-Ag-interacting hypo-Polα population and, accordingly, SV40 replication efficiency. Moreover, the ATR-Δp53 pathway facilitates the proteasomal degradation of the 180-kDa catalytic subunit of the non-T-Ag-interacting P-Polα, giving rise to T-Ag-interacting hypo-Polα. Thus, the purpose of activating the ATR-Δp53-p21-mediated intra-S checkpoint is to maintain the host in S phase, an optimal environment for SV40 replication, and to modulate the host DNA replicase, which is indispensable for viral amplification.Infection of quiescent CV-1 cells with the primate polyomavirus simian virus 40 (SV40) induces cell cycle progression and stimulates host cell DNA replication, which is mandatory for viral amplification. SV40 uses only a single viral protein, T antigen (T-Ag), for its own replication; all other components have to be provided by the host. Initially, a specifically phosphorylated subclass of T-Ag binds to a palindromic sequence in the SV40 origin (43), and in the presence of ATP, T-Ag forms a double-hexamer nucleoprotein complex leading to structural distortion and unwinding of origin DNA sequences (5). In concert with the cellular single-strand DNA binding protein RPA and topoisomerase I, the DNA helicase activity of T-Ag promotes more-extensive origin unwinding, forming a preinitiation complex (pre-RC), resulting in an initiation complex (53). Once the initiation complex forms, the primase activity of the heterotetrameric DNA polymerase α-primase (Polα) complex, consisting of the p180 catalytic subunit, the p70 regulatory subunit, and the p48/58 primase subunits, synthesizes a short RNA primer on each template strand, which is extended by the DNA polymerase activity of Polα (6, 17). Immediately after the first nascent RNA/DNA primer is synthesized, the complete replication machinery is assembled, and elongation at both forks by the processive DNA polymerase δ ensues (62). Thus, during the initiation of SV40 replication, T-Ag performs many of the functions attributed to the eukaryotic pre-RC complex proteins, including Orc, Cdc6, Cdt1, and kinase-independent cyclin E, which facilitates loading of the putative replication helicase Mcm2-7 onto the eukaryotic origin (4, 18). Biochemical evidence shows that initiation of SV40 and eukaryotic DNA replication occurs by the physical interaction of Polα with the appropriate pre-RC in the immediate vicinity of the origin. In SV40, Polα is loaded onto the origin by direct physical contact between the helicase T-Ag and its p180 N-terminal domain C (14, 15, 16). In eukaryotes, Cdc45, Mcm10, and And-1 cooperate to recruit Polα to the origin-initiation complex, thereby tethering the replicase to the origin-loaded Mcm2-7 helicase (34, 61).Although SV40 and chromosomal DNA replication share the same essential replication factors that are recruited to the appropriate pre-RC, there are noticeable differences between the SV40 and eukaryotic replication systems. The viral system allows unregulated multiple firing of the origin, whereas in the eukaryotic system, origin-dependent initiation of replication is regulated and restricted to firing only once per cell cycle. Reinitiation at origins within a cell cycle is prevented by the inactivation of pre-RC components in S and G2. The cyclin-dependent kinases (Cdks) play a central role in establishing a block to rereplication through phosphorylation of each of the components. At present, several proteins of the mammalian pre-RC, such as Orc1, Cdt1, Cdc6, and the Mcm complex are phosphorylated by cyclin A (cycA)-Cdk2/1 (AK) and, as a result, are degraded or inactivated (1, 26, 30, 33, 40). Nevertheless, not all of the pre-RC components mentioned above are utilized by SV40, and accordingly, not all are involved in viral initiation control. However, in both replication systems, DNA synthesis is initiated by Polα and its initiation activity is regulated by Cdks (55). Moreover, AK-phosphorylated Polα is not recruited to mammalian origins in vivo (13) and is unable to initiate SV40 replication in vitro (47, 57, 58). Considering that cellular mechanisms blocking the rereplication of DNA act by AK phosphorylation of the replication factors mentioned above, especially Polα, it is feasible to suggest that downregulation or even inhibition of this kinase activity promotes dysregulation of replication control in SV40-infected cells.One pathway that leads to downregulation of AK activity in response to cellular stress is the intra-S checkpoint, which employs the novel p53 isoform Δp53 (45). In UV-damaged S-phase cells, ATR (ataxia-telangiectasia mutated [ATM]-Rad3 related)-activated Δp53 upregulates the Cdk inhibitor p21, resulting in a transient reduction in AK activity and decelerated S-phase progression (45). Here we demonstrate that SV40 lytic infection triggers the ATR signaling cascade, leading to the activation of Δp53 and accordingly a p21-mediated drop in AK activity to prevent AK-catalyzed inactivation of the hypophosphorylated, T-Ag-interacting Polα (hypo-Polα) subclass, which is essential for the initiation of viral DNA replication.  相似文献   

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.
The molecular complexes involved in the nonhomologous end-joining process that resolves recombination-activating gene (RAG)-induced double-strand breaks and results in V(D)J gene rearrangements vary during mammalian ontogeny. In the mouse, the first immunoglobulin gene rearrangements emerge during midgestation periods, but their repertoires have not been analyzed in detail. We decided to study the postgastrulation DJH joints and compare them with those present in later life. The embryo DJH joints differed from those observed in perinatal life by the presence of short stretches of nontemplated (N) nucleotides. Whereas most adult N nucleotides are introduced by terminal deoxynucleotidyl transferase (TdT), the embryo N nucleotides were due to the activity of the homologous DNA polymerase μ (Polμ), which was widely expressed in the early ontogeny, as shown by analysis of Polμ−/− embryos. Based on its DNA-dependent polymerization ability, which TdT lacks, Polμ also filled in small sequence gaps at the coding ends and contributed to the ligation of highly processed ends, frequently found in the embryo, by pairing to internal microhomology sites. These findings show that Polμ participates in the repair of early-embryo, RAG-induced double-strand breaks and subsequently may contribute to preserve the genomic stability and cellular homeostasis of lymphohematopoietic precursors during development.The adaptive immune system is characterized by the great diversity of its antigen receptors, which result from the activities of enzymatic complexes that cut and paste the genomic DNA of antigen receptor loci. The nonhomologous end-joining (NHEJ) machinery is then recruited to repair the double-strand DNA breaks (DSBs) inflicted by the products of the recombination-activating genes (RAGs) (45, 65). Within B cells, each immunoglobulin (Ig) receptor represents a singular shuffling of two heavy (H) and two light (L) chains, which are derived from the recombination of V, D, and J gene segments of the IgH locus and of V and J for IgL (71). Besides these combinatorial possibilities, most Ig variability derives from extensive processing of the coding ends, including exonucleolytic trimming of DNA ends, together with the addition of palindromic (P) nucleotides templated by the adjacent germ line sequence and of nontemplated (N) nucleotides secondary to the activity of the terminal deoxynucleotidyl transferase (TdT), a lymphoid-specific member of family X of DNA polymerases (reviewed in reference 56). During B-lineage differentiation, IgH rearrangements occur before those of the IgL locus, and D-to-JH rearrangements precede V-to-DJH rearrangements (62). DJH joints are formed in any of the three open reading frames (ORFs). ORF1 is predominantly used in mature Igs, ORF2 is transcribed as a Dμ protein that provides negative signals to the B-cell precursors, and ORF3 frequently leads to stop codons (32, 33, 37). Germ line V, D, and J gene segments display short stretches of mutually homologous nucleotides (SSH), which are frequently used in gene rearrangements during perinatal periods, when N additions are absent (27, 32, 55, 57). The actual Ig V-region repertoires represent both the results of the NHEJ process associated with genomic VDJ recombination and those of antigen-independent and -dependent selection events. Although the core NHEJ components (Ku-Artemis-DNA-PK and XLF-XRCC4-DNA ligase IV) are by themselves able to join RAG-induced, incompatible DNA ends, family X DNA polymerases can be recruited to fill gaps created by imprecise coding ends with 3′ overhangs (DNA polymerase μ [Polμ] and Polλ) and/or to promote diversity through the addition of N nucleotides (TdT) (34, 56).The lymphoid differentiation pathways and clonotypic repertoires are developmentally regulated and differ between the embryo-fetal and adult periods (2, 44, 68). The perinatal B cells result from a wave of B lymphopoiesis occurring during the last third of mouse gestation (13, 14, 21, 70). Perinatal VH gene usage differs from that predominating in the adult (1, 69), and the former VDJ joints rarely display N additions, leading to V-region repertoires enriched in multi- and self-reactive specificities (36, 40). The program of B-cell differentiation starts at embryonic days 10 to 11 (E10 to E11) in embryo hematopoietic sites, after the emergence of multipotent progenitors (at E8.5 to E9.5) (18, 19, 23, 31, 51, 73). DJH rearrangements were detected in these early embryos, whereas full VDJH sequences were not observed before E14 (14, 18, 51, 66), when VJκ rearrangements were also found (63). The earliest mouse DJH/VDJH Ig sequences analyzed to date corresponded to late fetuses (E16) (14, 53). We reasoned that the true baseline of the Ig rearrangement process occurs in midgestation embryos, when the first DJHs are not yet transcribed and, consequently, not subjected to selection and are conditioned only for the evolutionarily established and developmentally regulated usage of distinct NHEJ machineries.We report here the sequence profiles of the earliest embryo E10 to E12 DJH joints. Unexpected frequencies of embryonic DJH joints bearing N nucleotides, in the absence of detectable TdT expression, were found. Moreover, the embryo DJH joints lacking N nucleotides (N) used fewer SSH to recombine than newborn DJHs, and these SSH were widely dispersed along the embryo D sequences, in contrast to the most joint-proximal ones, which predominated in newborn DJHs. Considering that Polμ is the closest relative of TdT (42% amino acid identity) (22), which is able to introduce N nucleotides in vitro (4, 22, 34, 39, 49) and to join DNA ends with minimal or even null complementarity (17, 58), and that it is expressed in early-embryo organs, we decided to investigate its putative contribution to the first embryo DJH joints. The DJH joints obtained from Polμ−/− embryos (48) showed a significant reduction of N nucleotides compared to wild-type (WT) embryos. Moreover, highly preserved DJH joints (with <3 deleted nucleotides) were selectively depleted in the Polμ−/− mouse embryos, while the remaining DJHs preferentially relied upon longer stretches of homology for end ligation. These findings support the idea that Polμ is active during early-embryo DJH rearrangements and that both its template-dependent and -independent ambivalent functions may be used to fill in small nucleotide gaps generated after asymmetric hairpin nicking and also to extend coding ends via a limited TdT-like activity.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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