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.
3.
4.
One essential downstream signaling pathway of receptor tyrosine kinases (RTKs), such as vascular endothelial growth factor receptor (VEGFR) and the Tie2 receptor, is the phosphoinositide-3 kinase (PI3K)-phosphoinositide-dependent protein kinase 1 (PDK1)-Akt/protein kinase B (PKB) cascade that plays a critical role in development and tumorigenesis. However, the role of PDK1 in cardiovascular development remains unknown. Here, we deleted PDK1 specifically in endothelial cells in mice. These mice displayed hemorrhage and hydropericardium and died at approximately embryonic day 11.5 (E11.5). Histological analysis revealed defective vascular remodeling and development and disrupted integrity between the endothelium and trabeculae/myocardium in the heart. The atrioventricular canal (AVC) cushion and valves failed to form, indicating a defect in epithelial-mesenchymal transition (EMT), together with increased endothelial apoptosis. Consistently, ex vivo AVC explant culture showed impeded mesenchymal outgrowth. Snail protein was reduced and was absent from the nucleus in AVC cells. Delivery of the Snail S6A mutant to the AVC explant effectively rescued EMT defects. Furthermore, adenoviral Akt delivery rescued EMT defects in AVC explant culture, and deletion of PTEN delayed embryonic lethality of PDK1 endothelial deletion mice by 1 day and rendered normal development of the AVC cushion in the PDK1-deficient heart. Taken together, these results have revealed an essential role of PDK1 in cardiovascular development through activation of Akt and Snail.Polypeptide growth factors, such as insulin, insulin-like growth factor 1 (IGF-I), vascular endothelial growth factor (VEGF), and angiopoietin 1 (Ang1), exert biological functions through binding to their transmembrane receptors that belong to a large family of receptor tyrosine kinases (RTKs) (4). Consequently, the receptor molecules form homo- or heterodimers, and the intracellular tyrosines at the carboxyl termini of the receptors become phosphorylated (37). Numerous distinct adaptor/regulatory proteins, through their Src homologous 2 (SH2) domains, bind to the phosphotyrosines and transduce the signal to downstream pathways, among which are two essential and well-characterized signaling cascades—the mitogen-activated protein kinase (MAPK) and phosphoinositide-3 kinase (PI3K)-phosphoinositide-dependent protein kinase 1 (PDK1)-Akt signaling pathways (4, 13, 37).The regulatory subunit of PI3K, p85, possesses the SH2 domain and can, therefore, bind to phosphotyrosines on the RTKs and subsequently render activation of the catalytic subunit of PI3K, p110 (7, 8). Active p110 phosphorylates phosphoinositide biphosphate (PIP2), turning it into PIP3 that recruits PDK1 and Akt to the cellular membrane, where Akt is phosphorylated at threonine 308 (T308 for Akt1) by PDK (5, 23, 30). The serine 473 (S473) of Akt (Akt1) is phosphorylated by mTOR complex 2 (mTORC2) and other kinases (17, 36). Phosphorylation of Akt at these two amino acids brings it to full activation. In PDK1-deficient embryonic stem (ES) cells, T308 phosphorylation was abolished and most of the Akt activity was lost, although the S473 phosphorylation was intact (40).Akt plays an important role in multiple biological processes, such as cell survival, growth, glucose metabolism, and angiogenesis (2, 12, 14-16, 22, 23, 39, 41-43). In mammals, there are three Akt isoforms, termed Akt 1, -2, and -3. Previously, we generated Akt1- and Akt3-deficient mice and studied their roles in mouse development (2, 15, 39, 42, 43). We found that the Akt1 and -3 double knockout (KO) (DKO) mice were embryonically lethal at around embryonic day 12 (E12) and manifested developmental defects in multiple tissues, including the cardiovascular system (14, 15, 43). These studies suggest that the Akt signaling pathway is involved in cardiovascular development.Other than Akt isoforms, PDK1 also activates another group of AGC family kinases, such as p70 ribosomal S6 kinase (S6K) (32), serum, and glucocorticoid-induced protein kinase (SGK) (26), p90 ribosomal S6 kinase (RSK) (21), and atypical isoforms of protein kinase C (PKC) (31). Comprehensive and intensive mouse genetic studies performed mainly by Alessi and coworkers have confirmed the regulation of these AGC kinases by PDK1 (3, 9, 10, 27-29, 40).PDK1 knockout mice were severely growth retarded and died at around E9.0, indicating an essential role of PDK1 in development (27). However, its function and downstream targets in cardiovascular development are still elusive. To study this, we deleted PDK1 specifically in endothelial cells through Cre recombinase-mediated excision (25). The results have revealed an essential role of PDK1 in vascular remodeling and integrity and in cardiac development through activation of Akt and its downstream target of Snail.  相似文献   

5.
6.
Protein tyrosine kinase 6 (PTK6) is a nonmyristoylated Src-related intracellular tyrosine kinase. Although not expressed in the normal mammary gland, PTK6 is expressed in a majority of human breast tumors examined, and it has been linked to ErbB receptor signaling and AKT activation. Here we demonstrate that AKT is a direct substrate of PTK6 and that AKT tyrosine residues 315 and 326 are phosphorylated by PTK6. Association of PTK6 with AKT occurs through the SH3 domain of PTK6 and is enhanced through SH2 domain-mediated interactions following tyrosine phosphorylation of AKT. Using Src, Yes, and Fyn null mouse embryonic fibroblasts (SYF cells), we show that PTK6 phosphorylates AKT in a Src family kinase-independent manner. Introduction of PTK6 into SYF cells sensitized these cells to physiological levels of epidermal growth factor (EGF) and increased AKT activation. Stable introduction of active PTK6 into SYF cells also resulted in increased proliferation. Knockdown of PTK6 in the BPH-1 human prostate epithelial cell line led to decreased AKT activation in response to EGF. Our data indicate that in addition to promoting growth factor receptor-mediated activation of AKT, PTK6 can directly activate AKT to promote oncogenic signaling.Protein tyrosine kinase 6 (PTK6; also known as the breast tumor kinase BRK) is an intracellular Src-related tyrosine kinase (9, 48). Human PTK6 was identified in cultured human melanocytes (32) and breast tumor cells (39), while its mouse orthologue was cloned from normal small intestinal epithelial cell RNA (50). Although PTK6 shares overall structural similarity with Src family tyrosine kinases, it lacks an N-terminal myristoylation consensus sequence for membrane targeting (39, 51). As a consequence, PTK6 is localized to different cellular compartments, including the nucleus (14, 15). PTK6 is expressed in normal differentiated epithelial cells of the gastrointestinal tract (34, 42, 51), prostate (14), and skin (51-53). Expression of PTK6 is upregulated in different types of cancers, including breast carcinomas (6, 39, 54), colon cancer (34), ovarian cancer (47), head and neck cancers (33), and metastatic melanoma cells (16). The significance of apparent opposing signaling roles for PTK6 in normal differentiation and cancer is still poorly understood.In human breast tumor cells, PTK6 enhances signaling from members of the ErbB receptor family (10, 29, 30, 36, 40, 49, 54). In the HB4a immortalized human mammary gland luminal epithelial cell line, PTK6 promoted epidermal growth factor (EGF)-induced ErbB3 tyrosine phosphorylation and AKT activation (29). In response to EGF stimulation, PTK6 promoted phosphorylation of the focal adhesion protein paxillin and Rac1-mediated cell migration (10). PTK6 can be activated by the ErbB3 ligand heregulin and promotes activation of extracellular signal-regulated kinase 5 (ERK5) and p38 mitogen-activated protein kinase (MAPK) in breast cancer cells (40). PTK6 can also phosphorylate p190RhoGAP-A and stimulate its activity, leading to RhoA inactivation and Ras activation and thereby promoting EGF-dependent breast cancer cell migration and proliferation (49). Expression of PTK6 has been correlated with ErbB2 expression in human breast cancers (4, 5, 54).AKT (also called protein kinase B) is a serine-threonine kinase that is activated downstream of growth factor receptors (38). It is a key player in signaling pathways that regulate energy metabolism, proliferation, and cell survival (7, 45). Aberrant activation of AKT through diverse mechanisms has been discovered in different cancers (2). AKT activation requires phosphorylation of AKT on threonine residue 308 and serine residue 473. The significance of phosphorylation of AKT on tyrosine residues is less well understood. Src has been shown to phosphorylate AKT on conserved tyrosine residues 315 and 326 near the activation loop (11). Substitution of these two tyrosine residues with phenylalanine abolished AKT kinase activity stimulated by EGF (11). Use of the Src family inhibitor PP2 impaired AKT activation following IGF-1 stimulation of oligodendrocytes (13). The RET/PTC receptor tyrosine kinase that responds to glial cell-line-derived neurotrophic factor also phosphorylated AKT tyrosine residue 315 promoting activation of AKT (28). AKT tyrosine residue 474 was phosphorylated when cells were treated with the tyrosine phosphatase inhibitor pervanadate, and phosphorylation of tyrosine 474 contributed to full activation of AKT (12). Recently, the nonreceptor tyrosine kinase Ack1 was shown to regulate AKT tyrosine phosphorylation and activation (37).Here we show that AKT is a cytoplasmic substrate of the intracellular tyrosine kinase PTK6. We identify the tyrosine residues on AKT that are targeted by PTK6, and we demonstrate that tyrosine phosphorylation plays a role in regulating association between PTK6 and AKT. In addition, we show that PTK6 promotes AKT activation and cell proliferation in a Src-independent manner.  相似文献   

7.
Oxidant stress is a ubiquitous stressor with negative impacts on multiple cell types. ASK1 is a central mediator of oxidant injury, but while mechanisms of its inhibition, such as sequestration by 14-3-3 proteins and thioredoxin, have been identified, mechanisms of activation have remained obscure and the signaling pathways regulating this are not clear. Here, we report that phosphorylation of 14-3-3ζ at serine 58 (S58) is dynamically regulated in the cell and that the phosphorylation status of S58 is a critical factor regulating oxidant stress-induced cell death. Phosphorylation of S58 releases ASK1 from 14-3-3ζ, and ASK1 then activates stress-activated protein kinases, leading to cell death. While several members of the mammalian sterile 20 (Mst) family of kinases can phosphorylate S58 when overexpressed, we identify Ste20/oxidant stress response kinase 1 (SOK-1), an Mst family member known to be activated by oxidant stress, as a central endogenous regulator of S58 phosphorylation and thereby of ASK1-mediated cell death. Our findings identify a novel pathway that regulates ASK1 activation and oxidant stress-induced cell death.Oxidant stress plays a central role in a wide variety of pathologies, and a critical mediator of oxidant injury is the protein kinase ASK1 (30). Indeed, ASK1 is required for several types of oxidant stress-induced cell death (32). Its activity is restrained by a large number of complementary mechanisms, a fact that attests to the importance of ASK1 being maintained in an inactive state in the cell. For example, reduced thioredoxin binds to the N-terminal region of ASK1, thereby inhibiting its activity (27). Following oxidant stress and oxidation of thioredoxin, ASK1 is released, allowing its activation. Multiple phosphorylation events, including phosphorylation of ASK1 at S83 by Akt and at S1033 by an unknown mechanism, also negatively regulate ASK1 (6, 41; reviewed in reference 30). Critical to the negative regulation of ASK1 is phosphorylation of S966, which drives the association of ASK1 with 14-3-3 proteins, thereby inhibiting ASK1-mediated activation of downstream signaling and cell death (8, 43). The kinases responsible for S966 phosphorylation are not known, but the protein phosphatase calcineurin has been shown to dephosphorylate S966, leading to dissociation of ASK1 from 14-3-3 (13). Thus, other than calcineurin-mediated dephosphorylation of ASK1, signaling mechanisms positively regulating the release of ASK1 from 14-3-3 proteins are not known, despite intense interest in this kinase as a potential target in cardiovascular and neurologic diseases (30). Therefore, we undertook studies to attempt to identify such a mechanism.14-3-3 proteins play protective roles in the cell by sequestering proapoptotic factors in a phosphorylation-dependent manner (1, 15, 23). These proapoptotic proteins that are sequestered by 14-3-3 proteins are typically phosphorylated on one or more 14-3-3 binding motifs (18, 39). For example, in addition to ASK1 phosphorylation at S966, (8, 30), Bad is phosphorylated by Akt and ribosomal S6 kinases at several residues, inhibiting its proapoptotic functions (4, 14, 42, 45). Acting in opposition to this is the well-characterized c-Jun N-terminal kinase (JNK)-mediated phosphorylation of serine 184 of 14-3-3 proteins, leading to release of the proapoptotic factors Bax, Bad, FOXO3a, and Abl (29, 33, 40). In addition to S184, the phosphorylation statuses of other 14-3-3 residues can regulate 14-3-3/client interactions, such as T233, which is phosphorylated by CKI, disrupting the 14-3-3/Raf-1 interaction (5).Although most of the attention to phosphorylation of 14-3-3 has been focused on S184 and T233 (1), S58 has been known to be phosphorylated in situ for some time, and several kinases have been implicated, including protein kinases A and D, Akt, mitogen-activated protein kinase-activated kinase 2 (MK2), and sphingosine-dependent protein kinase 1 (later identified as a cleavage fragment of protein kinase C δ) (9, 16, 17, 24, 25, 44). However, it is not clear which specific kinases mediate phosphorylation under specific circumstances, nor are the biological consequences clear. This is underscored by the fact that both pro- and antiapoptotic kinases have been reported to phosphorylate this residue (23). It does seem clear, however, that S58 phosphorylation disrupts 14-3-3 dimerization and that this reduces the binding of some proteins (e.g., Raf-1) (28, 34), though probably not all, since Woodcock et al. reported that 14-3-3ζ monomers phosphorylated at S58 remained competent to bind phosphopeptides (37).Thorson et al. and Wang et al. created 14-3-3 mutants that were deficient in binding phosphopeptides, and Xing et al. employed one of these, 14-3-3ζ(R56A/R60A), to show that it led to enhanced activation of the stress-activated protein kinases, JNKs and p38, and enhanced cell death in response to UVC irradiation, a model of oxidant stress (31, 36, 38). However, since S58 of 14-3-3ζ is in the center of the R56-R60 region, we hypothesized that phosphorylation of S58 might disrupt binding of 14-3-3ζ to ASK1, which is upstream from the JNKs and p38 in the response to oxidant stress. In addition, Preisinger et al. reported that Ste20/oxidant stress response kinase 1 (SOK-1)/STK25, a mammalian sterile 20 (Mst) family member also known as yeast sterile 20 kinase 1 (YSK1), phosphorylated 14-3-3ζ on S58, leading to reorganization of the Golgi structure and cell motility (26). However, we and others had initially identified SOK-1 as a kinase activated by oxidant stress (20-22), and we have also shown that SOK-1 exits the Golgi apparatus following cellular stress (19), potentially giving SOK-1 greater access to the ASK1/14-3-3 complex (which is primarily cytosolic). Thus, we hypothesized that oxidant stress-induced activation of SOK-1 might lead to S58 phosphorylation, release of ASK1 from 14-3-3ζ, activation of JNKs and p38, and cell death. Here, we confirm these hypotheses and identify S58 as a molecular switch regulating ASK1-mediated oxidant stress-induced cell death.  相似文献   

8.
9.
Us3 protein kinases encoded by herpes simplex virus 1 (HSV-1) and 2 (HSV-2) are serine/threonine protein kinases and play critical roles in viral replication and pathogenicity in vivo. In the present study, we investigated differences in the biological properties of HSV-1 and HSV-2 Us3 protein kinases and demonstrated that HSV-2 Us3 did not have some of the HSV-1 Us3 kinase functions, including control of nuclear egress of nucleocapsids, localization of UL31 and UL34, and cell surface expression of viral envelope glycoprotein B. In agreement with the observations that HSV-2 Us3 was less important for these functions, the effect of HSV-2 Us3 kinase activity on virulence in mice following intracerebral inoculation was much lower than that of HSV-1 Us3. Furthermore, we showed that alanine substitution in HSV-2 Us3 at a site (aspartic acid at position 147) corresponding to one that can be autophosphorylated in HSV-1 Us3 abolished HSV-2 Us3 kinase activity. Thus, the regulatory and functional effects of Us3 kinase activity are different between HSV-1 and HSV-2.Us3 protein kinases encoded by herpes simplex virus 1 (HSV-1) and 2 (HSV-2) are serine/threonine protein kinases with amino acid sequences that are conserved in the subfamily Alphaherpesvirinae (6, 24, 36). Based on studies showing that recombinant Us3 mutants of HSV-1 and HSV-2 have significantly impaired viral replication and virulence in mice models, it has been concluded that both HSV-1 and HSV-2 Us3 protein kinases play important roles in viral replication and pathogenicity in vivo (25, 33, 41). In contrast, HSV-1 and HSV-2 Us3 protein kinases are not essential for growth in tissue culture cells (33, 36). Thus, recombinant Us3 mutants grow as well as wild-type viruses in Vero cells, and the mutants exhibit modestly impaired replication in HEp-2 cells (33, 36, 39, 40). The possible functions of Us3 have been extensively studied and gradually elucidated for HSV-1 Us3, but much less is known about HSV-2 Us3. These functions include (i) blocking apoptosis (1, 22, 30, 31, 35); (ii) promoting nuclear egress of progeny nucleocapsids through the nuclear membrane (39, 40, 45); (iii) redistributing and phosphorylating nuclear membrane-associated viral nuclear egress factors UL31 and UL34 (14, 37, 38) and cellular proteins, including lamin A/C and emerin (21, 27, 28); (iv) controlling infected cell morphology (13, 31, 32); and (v) downregulating cell surface expression of viral envelope glycoprotein B (gB) (12).To determine the molecular mechanisms for a viral protein kinase''s effects in infected cells, the kinase''s physiological substrates and its phosphorylation sites must be identified. This can involve studies showing that the altered phenotypes observed in cells infected with a mutant virus lacking the protein kinase activity is also detected in cells infected with a mutant virus in which the substrate''s phosphorylation sites have been modified by mutations. Although more than 15 potential HSV Us3 substrates have been reported, HSV-1 Us3 phosphorylation of only three substrates (Us3 itself, gB, and UL31) has been demonstrated to be linked directly with Us3 functions in infected cells (12, 13, 29, 41) as follows. (i) Us3 has been reported to autophosphorylate serine at position 147 (Ser-147), and this phosphorylation augments Us3''s kinase activity in infected cells (13, 41). Even though only a small fraction of Us3 is autophosphorylated at Ser-147 in infected cells, alanine replacement of Ser-147 in Us3 significantly reduced HSV-1 replication in the mouse cornea and pathogenic manifestations of herpes stroma keratitis and periocular skin disease in mice (41). These results indicated that Us3 kinase activity was, in part, regulated by autophosphorylation of Ser-147, and regulation of Us3 activity by autophosphorylation played a critical role in viral replication in vivo and HSV-1 pathogenesis. (ii) It has been reported that HSV-1 Us3 phosphorylates Thr-887 in the cytoplasmic tail of gB, and this phosphorylation downregulates the cell surface expression of gB (12). Us3 phosphorylation of gB at Thr-887 also has been proposed to be involved in the regulation of fusion of the nascent progeny virion envelope with the cell''s outer nuclear membrane, based on the observation that virions accumulated aberrantly in the perinuclear space in cells infected with mutant viruses carrying the amino acid substitution mutation T887A in gB and lacking the capacity to produce gH (45). The Us3 phosphorylation of gB at Thr-887 appeared to be critical for HSV-1 replication and pathogenesis in vivo, based on studies showing that the T887A substitution in the phosphorylation site in gB significantly reduced viral replication in the mouse cornea and pathogenic manifestations of herpes stroma keratitis and periocular skin disease in mice (Takahiko Imai, Ken Sagou, and Yasushi Kawaguchi, unpublished observations). (iii) It has been shown that Us3 phosphorylated some or all of the six serines in the UL31 N-terminal region, and this phosphorylation regulated the proper localization of UL31 and UL34 at the nuclear membrane and nuclear egress of nucleocapsids (29). Thus, the molecular basis of HSV-1 Us3 effects in infected cells have been gradually elucidated.However, the Us3 phosphorylation sites in Us3 itself and in gB are not conserved between HSV serotypes (12, 13). The amino acid residues in HSV-2 Us3 and gB corresponding to HSV-1 Us3 Ser-147 and gB Thr-887 are aspartic acid (Asp-147) and alanine (Ala-887), respectively. These results suggest that some HSV-1 Us3 functions, such as regulation of its own catalytic activity and control of gB expression on the cell surface, are not regulated by HSV-2 Us3 or are regulated in a manner(s) different from HSV-1 Us3. In agreement with this suggestion, there is a marked difference between HSV-1 and HSV-2 virulence in mice following intracerebral infection, with the HSV-1 Us3 null mutant being >104-fold less virulent than the parent wild-type virus (25), while the HSV-2 Us3 null mutant was only ∼10-fold less virulent (33). Although these results were from different reports and the mouse strains used in the studies were different, they indicate that some HSV-1 Us3 functions are different from those of HSV-2 Us3.Therefore, we investigated differences in the biological properties of HSV-1 and HSV-2 Us3 protein kinases. It was of particular interest to examine whether Asp-147 in HSV-2 Us3 is required for its own kinase activity, since it is well established that acidic amino acids such as Asp or glutamic acid sometimes mimic the negative charges produced by phosphorylation (29, 46). In the present study, using a genetic manipulation system of HSV-2 with our newly constructed HSV-2 bacterial artificial chromosome (BAC) clone, we have shown that HSV-2 Us3 exhibited marked differences from HSV-1 Us3 in its catalytic functions, including the regulation of UL31/UL34 localization, nuclear egress of nucleocapsids, cell surface expression of gB, and virulence in mice. We also found that Asp-147 in HSV-2 Us3 was critical for its kinase activity, raising a possibility that the activity of Us3 kinases was regulated differently in HSV-1 and HSV-2.  相似文献   

10.
KSR1 is a mitogen-activated protein (MAP) kinase scaffold that enhances the activation of the MAP kinase extracellular signal-regulated kinase (ERK). The function of KSR1 in NK cell function is not known. Here we show that KSR1 is required for efficient NK-mediated cytolysis and polarization of cytolytic granules. Single-cell analysis showed that ERK is activated in an all-or-none fashion in both wild-type and KSR1-deficient cells. In the absence of KSR1, however, the efficiency of ERK activation is attenuated. Imaging studies showed that KSR1 is recruited to the immunological synapse during T-cell activation and that membrane recruitment of KSR1 is required for recruitment of active ERK to the synapse.Kinase suppressor of Ras was originally identified in Drosophila melanogaster (53) and Caenorhabditis elegans (19, 32, 52) as a positive regulator of the extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase signaling pathway. It is thought to function as a MAP kinase scaffold because it can bind to Raf, MEK, and ERK (18, 19, 27, 28, 44, 59). While the exact function of KSR is unknown, preassembling the three components of the ERK MAP kinase cascade could function to enhance the efficiency of ERK activation, potentially regulate the subcellular location of ERK activation, and promote access to specific subcellular substrates (16, 45, 46).While only one isoform of KSR is expressed in Drosophila (53), two KSR isoforms have been identified in C. elegans (19, 32, 52) and most higher organisms. They are referred to as KSR1 and KSR2 (32, 43). While KSR1 mRNA and protein are detectable in a wide variety of cells and tissues, including brain, thymus, and muscle (10, 11, 29), little is known about the expression pattern of KSR2.We previously reported the phenotype of KSR1-deficient mice (30). These mice are born at Mendelian ratios and develop without any obvious defects. Using gel filtration, we showed that KSR1 promotes the formation of large signaling complexes containing KSR1, Raf, MEK, and ERK (30). Using both primary T cells stimulated with antibodies to the T-cell receptor as well as fibroblasts stimulated with growth factors, we showed that KSR1-deficient cells exhibit an attenuation of ERK activation with defects in cell proliferation.Here we explored the role of KSR1 in NK cell-mediated cytolysis. The killing of a target cell by a cytolytic T cell or NK cell is a complicated process that involves cell polarization with microtubule-dependent movement of cytolytic granules to an area that is proximal to the contact surface or immunological synapse (7, 33, 34, 48-50, 54). A variety of different signaling molecules are also involved, including calcium (23), phosphatidylinositol-3,4,5-triphosphate (13, 17), and activation of the ERK MAP kinase (6, 42, 56). Recently, the recruitment of activated ERK to the immunological synapse (IS) has been shown to be a feature of successful killing of a target by cytotoxic T lymphocytes (58).How active ERK is recruited to the synapse is not known. Since KSR1 is known to be recruited to the plasma membrane by Ras activation (24), and since the immunological synapse is one of the major sites of Ras activation (26, 41), it seemed plausible to test the hypothesis that KSR1 recruitment to the plasma membrane functions to recruit ERK to the immunological synapse and facilitate its activation. We found that KSR1 was recruited to the immunological synapse and that KSR1 appeared to be required for the localization of active ERK at the contact site. As KSR1-deficient cells exhibit a defect in killing, this suggests that KSR1 recruitment to the synapse may be important in the cytolytic killing of target cells.  相似文献   

11.
The hypoxia-regulated alternative TrkAIII splice variant expressed by human neuroblastomas exhibits oncogenic potential, driven by in-frame exon 6 and 7 alternative splicing, leading to omission of the receptor extracellular immunoglobulin C1 domain and several N-glycosylation sites. Here, we show that the TrkAIII oncogene promotes genetic instability by interacting with and exhibiting catalytic activity at the centrosome. This function depends upon intracellular TrkAIII accumulation and spontaneous interphase-restricted activation, in cytoplasmic tyrosine kinase (tk) domain orientation, predominantly within structures that closely associate with the fully assembled endoplasmic reticulum intermediate compartment and Golgi network. This facilitates TrkAIII tk-mediated binding of γ-tubulin, which is regulated by endogenous protein tyrosine phosphatases and geldanamycin-sensitive interaction with Hsp90, paving the way for TrkAIII recruitment to the centrosome. At the centrosome, TrkAIII differentially phosphorylates several centrosome-associated components, increases centrosome interaction with polo kinase 4, and decreases centrosome interaction with separase, the net results of which are centrosome amplification and increased genetic instability. The data characterize TrkAIII as a novel internal membrane-associated centrosome kinase, unveiling an important alternative mechanism to “classical” cell surface oncogenic receptor tk signaling through which stress-regulated alternative TrkAIII splicing influences the oncogenic process.Alternative splicing is fundamental for differential protein expression from the same gene and not only increases the proteomic complexity of higher organisms (29) but is also involved in cancer pathogenesis, activating several oncogenes and inactivating several oncosuppressors (17).The neurotrophin receptor tropomyosin-related kinase A (TrkA) is among the proto-oncogenes activated by alternative splicing, with a novel hypoxia-regulated oncogenic alternative TrkAIII splice variant recently identified in advanced-stage human neuroblastomas (NB) and primary glioblastomas (44, 45). In contrast to wild-type TrkAI/TrkAII, the expression of which is associated with better prognosis for NB, induces NB cell differentiation, exhibits a tumor suppressor function in NB models in vivo (9, 19, 22, 30, 44, 45), and may regulate both spontaneous and therapy-induced NB regression (30), TrkAIII is expressed by more-advanced-stage NB and exhibits oncogenic activity in NB models (44, 45). This has challenged the hypothesis of an exclusively tumor-suppressing function for TrkA in NB by providing a way through which tumor-suppressing signals from TrkA can be converted to oncogenic signals from TrkAIII during tumor progression.The oncogenic potential of TrkAIII, characterized by NIH 3T3 cell transforming and NB xenograft primary and metastatic tumor-promoting activity (44, 45), is driven by in-frame alternative splicing of exons 6 and 7. This results in the omission of the receptor extracellular immunoglobulin C1 (Ig C1) Ig-like domain and several N-glycosylation sites important in regulating TrkA cell surface expression and preventing ligand-independent activation (2, 44, 45, 48). As a consequence, and unlike TrkAI and TrkAII, TrkAIII is not expressed at the cell surface but accumulates in the intracellular membrane compartment, within which it exhibits spontaneous tyrosine kinase (tk) and phosphoinositol-3 kinase (PI3K) activity and induces chronic signaling through PI3K/Akt/NF-κB but not Ras/mitogen-activated protein kinase (MAPK), inducing a more stress-resistant, angiogenic, and tumorigenic NB cell phenotype (44, 45). This differs from ligand-activated cell surface TrkA, which signals transiently through Ras/MAPK in NB cells to induce differentiation and a less angiogenic and tumorigenic NB cell phenotype (9, 19, 22, 30, 44, 45). This difference in signaling provides a potential basis for the opposing tumor-suppressing and oncogenic effects of alternative TrkA splice variants, which may not only depend upon the dislocation of TrkAIII from cell surface caveolae, which are the sites of TrkAI expression and Ras/MAPK signal initiation (45, 48), but also TrkAIII-associated PI3K activity below the Ras/MAPK activation threshold and/or TrkAIII-associated PI3K antagonism of Raf/MEK/extracellular signal-regulated kinase signaling (44). Signal transduction from intracellular TrkAIII bears close resemblance to the transient signaling through PI3K/Akt/NF-κB but not Ras/MAPK induced by A2a adenosine receptor/c-Src-mediated transactivation of immature TrkAI within the Golgi network (GN) (37), suggesting that the intracellular localization of TrkAIII is a critical determinant of both differential signaling and oncogenic potential.Intracellular nonnuclear membranes are separated into the endoplasmic reticulum (ER), ER-GN intermediate compartment (ERGIC), GN, and transport vesicles, which assemble around, integrate, and interact with the centrosome (38). The centrosome, comprised of two centrioles embedded within a pericentriolar matrix of over 100 proteins, including γ-tubulin, acts as the major microtubule-organizing center and orchestrates the assembly, organization, and integration of the ER, ERGIC, GN, and associated vesicles (38). The centrosome also maintains genomic integrity by duplicating once per cell cycle S phase, ensuring bipolar mitotic spindle formation, accurate chromosome segregation, and the inheritance of a single centrosome by each daughter cell (23).Centrosome duplication is tightly regulated by protein kinases Aurora-A and -B, polo kinases 1 and 4 (Plk-1 and Plk-4), Cdk2, PI3K, Zyg-1, Syk, Nek2, regulators Pin-1 and separase, and related phosphatases (11, 12, 18, 31, 42, 47, 53). The deregulation of centrosome duplication leads to centrosome amplification and subsequently to aberrant mitotic spindle formation, which promotes aneuploidy and polyploidy. These manifestations of genetic instability represent hallmarks of malignancy and drive tumor progression by promoting a more malignant phenotype (11, 12, 26, 35, 49). Centrosome amplification and subsequent genetic instability are induced by kinases that target the centrosome, the loss of centrosome-associated kinase inhibitors, altered levels of centrosome-associated regulators, and oncosuppressor inactivation (6, 7, 11, 12, 35, 40, 42, 49, 53).The localization of TrkAIII to internal membranes, as a prerequisite for oncogenic activity (44, 45), makes identification of the membrane context within which TrkAIII exhibits activity and novel substrate interactions within this environment important in elucidating how TrkAIII exerts oncogenic potential. In the present study, we unveil a novel oncogenic mechanism for TrkAIII by demonstrating that TrkAIII activation within membranes that associate closely with the assembled ER/ERGIC/GN facilitates recruitment to the centrosome, results in centrosome amplification, and promotes genetic instability.  相似文献   

12.
The mammalian target of rapamycin (mTOR) is a conserved Ser/Thr kinase that forms two functionally distinct complexes important for nutrient and growth factor signaling. While mTOR complex 1 (mTORC1) regulates mRNA translation and ribosome biogenesis, mTORC2 plays an important role in the phosphorylation and subsequent activation of Akt. Interestingly, mTORC1 negatively regulates Akt activation, but whether mTORC1 signaling directly targets mTORC2 remains unknown. Here we show that growth factors promote the phosphorylation of Rictor (rapamycin-insensitive companion of mTOR), an essential subunit of mTORC2. We found that Rictor phosphorylation requires mTORC1 activity and, more specifically, the p70 ribosomal S6 kinase 1 (S6K1). We identified several phosphorylation sites in Rictor and found that Thr1135 is directly phosphorylated by S6K1 in vitro and in vivo, in a rapamycin-sensitive manner. Phosphorylation of Rictor on Thr1135 did not affect mTORC2 assembly, kinase activity, or cellular localization. However, cells expressing a Rictor T1135A mutant were found to have increased mTORC2-dependent phosphorylation of Akt. In addition, phosphorylation of the Akt substrates FoxO1/3a and glycogen synthase kinase 3α/β (GSK3α/β) was found to be increased in these cells, indicating that S6K1-mediated phosphorylation of Rictor inhibits mTORC2 and Akt signaling. Together, our results uncover a new regulatory link between the two mTOR complexes, whereby Rictor integrates mTORC1-dependent signaling.The mammalian target of rapamycin (mTOR) is an evolutionarily conserved phosphatidylinositol 3-kinase (PI3K)-related Ser/Thr kinase that integrates signals from nutrients, energy sufficiency, and growth factors to regulate cell growth as well as organ and body size in a variety of organisms (reviewed in references 4, 38, 49, and 77). mTOR was discovered as the molecular target of rapamycin, an antifungal agent used clinically as an immunosuppressant and more recently as an anticancer drug (5, 20). Recent evidence indicates that deregulation of the mTOR pathway occurs in a majority of human cancers (12, 18, 25, 46), suggesting that rapamycin analogs may be potent antineoplastic therapeutic agents.mTOR forms two distinct multiprotein complexes, the rapamycin-sensitive and -insensitive mTOR complexes 1 and 2 (mTORC1 and mTORC2), respectively (6, 47). In cells, rapamycin interacts with FKBP12 and targets the FKBP12-rapamycin binding (FRB) domain of mTORC1, thereby inhibiting some of its function (13, 40, 66). mTORC1 is comprised of the mTOR catalytic subunit and four associated proteins, Raptor (regulatory associated protein of mTOR), mLST8 (mammalian lethal with sec13 protein 8), PRAS40 (proline-rich Akt substrate of 40 kDa), and Deptor (28, 43, 44, 47, 59, 73, 74). The small GTPase Rheb (Ras homolog enriched in brain) is a key upstream activator of mTORC1 that is negatively regulated by the tuberous sclerosis complex 1 (TSC1)/TSC2 GTPase-activating protein complex (reviewed in reference 35). mTORC1 is activated by PI3K and Ras signaling through direct phosphorylation and inactivation of TSC2 by Akt, extracellular signal-regulated kinase (ERK), and p90 ribosomal protein S6 kinase (RSK) (11, 37, 48, 53, 63). mTORC1 activity is also regulated at the level of Raptor. Whereas low cellular energy levels negatively regulate mTORC1 activity through phosphorylation of Raptor by AMP-activated protein kinase (AMPK) (27), growth signaling pathways activating the Ras/ERK pathway positively regulate mTORC1 activity through direct phosphorylation of Raptor by RSK (10). More recent evidence has also shown that mTOR itself positively regulates mTORC1 activity by directly phosphorylating Raptor at proline-directed sites (20a, 75). Countertransport of amino acids (55) and amino acid signaling through the Rag GTPases were also shown to regulate mTORC1 activity (45, 65). When activated, mTORC1 phosphorylates two main regulators of mRNA translation and ribosome biogenesis, the AGC (protein kinase A, G, and C) family kinase p70 ribosomal S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), and thus stimulates protein synthesis and cellular growth (50, 60).The second mTOR complex, mTORC2, is comprised of mTOR, Rictor (rapamycin-insensitive companion of mTOR), mSin1 (mammalian stress-activated mitogen-activated protein kinase-interacting protein 1), mLST8, PRR5 (proline-rich region 5), and Deptor (21, 39, 58, 59, 66, 76, 79). Rapamycin does not directly target and inhibit mTORC2, but long-term treatment with this drug was shown to correlate with mTORC2 disassembly and cytoplasmic accumulation of Rictor (21, 39, 62, 79). Whereas mTORC1 regulates hydrophobic motif phosphorylation of S6K1, mTORC2 has been shown to phosphorylate other members of the AGC family of kinases. Biochemical and genetic evidence has demonstrated that mTORC2 phosphorylates Akt at Ser473 (26, 39, 68, 70), thereby contributing to growth factor-mediated Akt activation (6, 7, 52). Deletion or knockdown of the mTORC2 components mTOR, Rictor, mSin1, and mLST8 has a dramatic effect on mTORC2 assembly and Akt phosphorylation at Ser473 (26, 39, 79). mTORC2 was also shown to regulate protein kinase Cα (PKCα) (26, 66) and, more recently, serum- and glucocorticoid-induced protein kinase 1 (SGK1) (4, 22). Recent evidence implicates mTORC2 in the regulation of Akt and PKCα phosphorylation at their turn motifs (19, 36), but whether mTOR directly phosphorylates these sites remains a subject of debate (4).Activation of mTORC1 has been shown to negatively regulate Akt phosphorylation in response to insulin or insulin-like growth factor 1 (IGF1) (reviewed in references 30 and 51). This negative regulation is particularly evident in cell culture models with defects in the TSC1/TSC2 complex, where mTORC1 and S6K1 are constitutively activated. Phosphorylation of insulin receptor substrate-1 (IRS-1) by mTORC1 (72) and its downstream target S6K1 has been shown to decrease its stability and lead to an inability of insulin or IGF1 to activate PI3K and Akt (29, 69). Although the mechanism is unknown, platelet-derived growth factor receptor β (PDGF-Rβ) has been found to be downregulated in TSC1- and TSC2-deficient murine embryonic fibroblasts (MEFs), contributing to a reduction of PI3K signaling (80). Interestingly, inhibition of Akt phosphorylation by mTORC1 has also been observed in the presence of growth factors other than IGF-1, insulin, or PDGF, suggesting that there are other mechanisms by which mTORC1 activation restricts Akt activity in cells (reviewed in references 6 and 31). Recent evidence demonstrates that rapamycin treatment causes a significant increase in Rictor electrophoretic mobility (2, 62), suggesting that phosphorylation of the mTORC2 subunit Rictor may be regulated by mTORC1 or downstream protein kinases.Herein, we demonstrate that Rictor is phosphorylated by S6K1 in response to mTORC1 activation. We demonstrate that Thr1135 is directly phosphorylated by S6K1 and found that a Rictor mutant lacking this phosphorylation site increases Akt phosphorylation induced by growth factor stimulation. Cells expressing the Rictor T1135A mutant were found to have increased Akt signaling to its substrates compared to Rictor wild-type- and T1135D mutant-expressing cells. Together, our results suggest that Rictor integrates mTORC1 signaling via its phosphorylation by S6K1, resulting in the inhibition of mTORC2 and Akt signaling.  相似文献   

13.
14.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

15.
16.
Golgi fragmentation is a process that is necessary to allow its redistribution into daughter cells during mitosis, a process controlled by serine-threonine kinases. This Golgi fragmentation is activated by MEK1 and Plk3. Plk3 is a kinase that is a downstream target in the Golgi fragmentation pathway induced by MEK1 or by nocodazole. In this work, we have identified that Plk3 and VRK1 are two consecutive steps in this signaling pathway. Plk3 interacts with VRK1, forming a stable complex detected by reciprocal immunoprecipitations and pull-down assays; VRK1 colocalizes with giantin in the Golgi apparatus, as Plk3 also does, forming clearly detectable granules. VRK1 does not phosphorylate Plk3, but Plk3 phosphorylates the C-terminal region of VRK1 in Ser342. VRK1 with substitutions in S342 is catalytically active but blocks Golgi fragmentation, indicating that its specific phosphorylation is necessary for this process. The induction of Golgi fragmentation by MEK1 and Plk3 can be inhibited by kinase-dead VRK1, the knockdown of VRK1 by siVRK1, kinase-dead Plk3, or PD98059, a MEK1 inhibitor. The Plk3-VRK1 kinase module might represent two consecutive steps of a signaling cascade that participates in the regulation of Golgi fragmentation.The Golgi apparatus in mammalian cells is formed by cistern stacks, tubules, and small vesicles, which undergo extensive and sequential fragmentation in mitosis (33). The reorganization of the Golgi apparatus, involving fragmentation, dispersal, and reassembly, is tightly regulated during mitosis (1, 27, 30), and reversible phosphorylation plays a critical role (1, 21), although the components and their sequential organization in the context of the initiation or execution of the signal required for Golgi fragmentation are only partially known.Many signaling pathways are composed of consecutive kinases. Characterization of new signaling pathways requires the identification of their components, the connections between them, and the order in which they are organized. Human VRK1 is a novel serine-threonine kinase that phosphorylates several proteins implicated in cellular responses to stress and DNA damage, such as p53 (5, 20, 40), c-Jun (31), and ATF2 (32), as well as proteins needed for nuclear envelope assembly required at the end of mitosis, such as Baf (25). In addition, VRK1 kinase activity is inhibited by interaction with RanGDP, and this inhibition is relieved by RanGTP, suggesting an asymmetric distribution of its activity within the nucleus and in mitosis (29). These properties suggest that the VRK1 gene plays a role in the regulation of cell cycle initiation and/or progression, consistent with its requirement for entry into the cell cycle, where it behaves as an immediate-early response gene like c-MYC and FOS (36). The loss of VRK1 by use of small interfering RNA (siRNA) induces an early G1 block, before cyclin D1 expression (36), which is accompanied by a reduction in the phospho-retinoblastoma level and an accumulation of cycle inhibitors, such as p27 (36), resulting in a stop in cell cycle progression (36, 40).Several kinases are implicated in the control of cell proliferation and in different mitotic checkpoints; among them are the polo-like kinase (Plk) family, which is a group composed of four proteins (14, 39, 46). One of them, Plk3, contributes as a mediator of DNA damage checkpoint responses, since its kinase activity increases after oxidative stress (43) and induction of DNA damage by ionizing radiomimetic drugs (45). Plk3 physically interacts with and phosphorylates p53 in Ser20, and this interaction increases in response to DNA damage and induces either cell cycle arrest or apoptosis (44) so that genetic stability can be maintained by the prevention of the accumulation of genetic damage. Furthermore, Plk3 interacts with Chk2 (2, 45), an important mediator of DNA damage responses (6, 16), and there is a functional connection between them since Plk3 phosphorylates Chk2 in Ser62 and Ser73, which are necessary for full Chk2 activation by ATM (4). In mitotic cells, Plk3 is localized associated with the spindle poles and mitotic spindles, and deregulated expression of Plk3 induces cell cycle arrest and apoptosis by the perturbation of microtubule integrity (41). In addition, Plk3 expression is induced after mitogenic stimulation, and it is required for mitotic (28) and S-phase (48) entry. Plk3 also regulates Cdc25C (3, 23, 26) and the NF-κB signaling pathway (19). VRK1 phosphorylates p53 in Thr18 (20, 40), a residue phosphorylated in response to taxol, an inhibitor of microtubule polymerization (34).There is a possibility that VRK1 and Plk3 might be connected in some way, since subpopulations of both VRK1 (37) and Plk3 (28) have been detected in the Golgi apparatus near the centrosome, where they colocalize with Golgi markers such as giantin or GM130 (33). Golgi fragmentation can be induced by MEK1 (1, 15), and this signal is partly mediated by Plk3 (28, 42). Moreover, Golgi fragmentation is a required step during mitosis, occurring late in the G2/M phase of the cell cycle (11), and MEK1 is implicated in the activation of this process (1, 15).The common biological aspects of VRK1 and Plk3 proteins and the association of VRK1 and Plk3 subpopulations in the Golgi apparatus led us to think that there might be a functional connection between these two kinases and thus that they might be components in a common signaling pathway. In this work, we explored the possible connection between VRK1 and Plk3 and determined if they were functionally related in a biological process, Golgi fragmentation, in which one of them, Plk3, is already known to participate. This work demonstrates that Plk3 and VRK1 are consecutive components in the signaling pathway that induces Golgi fragmentation in mitosis.  相似文献   

17.
The parvovirus adeno-associated virus (AAV) contains a small single-stranded DNA genome with inverted terminal repeats that form hairpin structures. In order to propagate, AAV relies on the cellular replication machinery together with functions supplied by coinfecting helper viruses such as adenovirus (Ad). Here, we examined the host cell response to AAV replication in the context of Ad or Ad helper proteins. We show that AAV and Ad coinfection activates a DNA damage response (DDR) that is distinct from that seen during Ad or AAV infection alone. The DDR was also triggered when AAV replicated in the presence of minimal Ad helper proteins. We detected autophosphorylation of the kinases ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) and signaling to downstream targets SMC1, Chk1, Chk2, H2AX, and XRCC4 and multiple sites on RPA32. The Mre11 complex was not required for activation of the DDR to AAV infection. Additionally, we found that DNA-PKcs was the primary mediator of damage signaling in response to AAV replication. Immunofluorescence revealed that some activated damage proteins were found in a pan-nuclear pattern (phosphorylated ATM, SMC1, and H2AX), while others such as DNA-PK components (DNA-PKcs, Ku70, and Ku86) and RPA32 accumulated at AAV replication centers. Although expression of the large viral Rep proteins contributed to some damage signaling, we observed that the full response required replication of the AAV genome. Our results demonstrate that AAV replication in the presence of Ad helper functions elicits a unique damage response controlled by DNA-PK.Replication of viral genomes produces a large amount of extrachromosomal DNA that may be recognized by the cellular DNA damage machinery. This is often accompanied by activation of DNA damage response (DDR) signaling pathways and recruitment of cellular repair proteins to sites of viral replication. Viruses therefore provide good model systems to study the recognition and response to DNA damage (reviewed in reference 48). The Mre11/Rad50/Nbs1 (MRN) complex functions as a sensor of chromosomal DNA double-strand breaks (DSBs) and is involved in activation of damage signaling (reviewed in reference 41). The MRN complex also localizes to DNA DSBs and is found at viral replication compartments during infection with a number of DNA viruses (6, 40, 47, 70, 75, 77, 87, 93). The phosphatidylinositol 3-kinase-like kinases (PIKKs) ataxia telangiectasia mutated (ATM), ATM and Rad3-related kinase (ATR), and the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) are involved in the signal transduction cascades activated by DNA damage (reviewed in references 43, 51, and 71). These kinases respond to distinct types of damage and regulate DSB repair during different phases of the cell cycle (5), either through nonhomologous end-joining (NHEJ) or homologous recombination pathways (reviewed in references 63, 81, and 86). The DNA-PK holoenzyme is composed of DNA-PKcs and two regulatory subunits, the Ku70 and Ku86 heterodimer. DNA-PK functions with XRCC4/DNA ligase IV to repair breaks during NHEJ, and works with Artemis to process DNA hairpin structures during VDJ recombination and during a subset of DNA DSB events (46, 50, 86). While the kinase activity of DNA-PKcs leads to phosphorylation of a large number of substrates in vitro as well as autophosphorylation of specific residues (reviewed in references 16 and 85), it is currently unclear how DNA-PKcs contributes to signaling in cells upon different types of damage.The adeno-associated virus (AAV) genome consists of a molecule of single-stranded DNA with inverted terminal repeats (ITRs) at both ends that form double-hairpin structures due to their palindromic sequences (reviewed in reference 52). The ITRs are important for replication and packaging of the viral genome and for integration into the host genome. Four viral Rep proteins (Rep78, Rep68, Rep52, and Rep40) are also required for replication and packaging of the AAV genome into virions assembled from the Cap proteins. Although the Rep and Cap genes are replaced in recombinant AAV vectors (rAAV) that retain only the ITRs flanking the gene of interest, these vectors can be replicated by providing Rep in trans (reviewed in reference 7). Productive AAV infection requires helper functions supplied by adenovirus (Ad) or other viruses such as herpes simplex virus (HSV) (reviewed in reference 27), together with components of the host cell DNA replication machinery (54, 55, 58). In the presence of helper viruses or minimal helper proteins from Ad or HSV, AAV replicates in the nucleus at centers where the viral DNA and Rep proteins accumulate (35, 76, 84, 89). Cellular and viral proteins involved in AAV replication, including replication protein A (RPA), Ad DNA-binding protein (DBP), and HSV ICP8, localize with Rep proteins at these viral centers (29, 33, 76).A number of published reports suggest associations between AAV and the cellular DNA damage machinery. For example, transduction by rAAV vectors is increased by genotoxic agents and DNA damaging treatments (1, 62, 91) although the mechanisms involved remain unclear. Additionally, the ATM kinase negatively regulates rAAV transduction (64, 92), and we have shown that the MRN complex poses a barrier to both rAAV transduction and wild-type AAV replication (11, 67). UV-inactivated AAV particles also appear to activate a DDR involving ATM and ATR kinases that perturbs cell cycle progression (39, 60, 88). It has been suggested that this response is provoked by the AAV ITRs (60) and that UV-treated particles mimic stalled replication forks in infected cells (39). In addition to AAV genome components, the viral Rep proteins have been observed to exhibit cytotoxicity and induce S-phase arrest (3, 65).The role of cellular repair proteins in AAV genome processing has also been explored by examining the molecular fate of rAAV vectors, which are converted into circular and concatemeric forms that persist episomally (18, 19, 66). Proteins shown to regulate circularization in cell culture include ATM and the MRN complex (14, 64), while in vivo experiments using mouse models have implicated ATM and DNA-PK in this process (14, 20, 72). Additionally, DNA-PKcs and Artemis have recently been shown to cleave the ITR hairpins of rAAV vectors in vivo in a tissue-dependent manner (36). Despite these studies, it is not clear how damage response factors function together and how they impact AAV transduction and replication in human cells.In this study we examined the cellular response to AAV replication in the context of Ad infection or helper proteins. We show that coinfection with AAV and Ad activates a DDR that is distinct from that seen during infection with Ad alone. The ATM and DNA-PKcs damage kinases are activated and signal to downstream substrates, but the response does not require the MRN complex and is primarily mediated by DNA-PKcs. Although expression of the large Rep proteins induced some DDR events, full signaling appeared to require AAV replication and was accompanied by accumulation of DNA-PK at viral replication compartments. Our results demonstrate that AAV replication induces a unique DNA damage signal transduction response and provides a model system for studying DNA-PK.  相似文献   

18.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

19.
20.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  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