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 609 毫秒
1.
NOXO1 (Nox organizing protein 1) and NOXA1 (Nox activating protein 1) are homologs of p47phox and p67phox. p47phox functions in phagocytes as an essential organizing protein mediating the binding of other regulatory proteins during activation of the phagocyte oxidase, and its translocation to the membrane is triggered upon cell activation by hyperphosphorylation, which relieves autoinhibition of SH3 and PX domains. NOXO1 lacks an autoinhibitory region and phosphorylation sites that are present in p47phox. Co-transfection of Nox1, NOXO1, and NOXA1 reconstitutes ROS (reactive oxygen species) generation in HEK293 cells in the absence of cell stimulation. NOXO1 binds to the phosphatidylinositol (PtdIns) lipids PtdIns 3,5-P2, PtdIns 5-P, and PtdIns 4-P. Unlike p47phox, which is located in the cytosol of resting cells and translocates to the plasma membrane where gp91phox is located, NOXO1 co-localizes with Nox1 in the membranes of resting cells. This localization of NOXO1 is dictated by its PX domain, since this domain but not the remainder of the molecule localizes to membranes. A point mutation in the PX domain of holo-NOXO1 decreases lipid binding resulting in cytosolic localization and also inhibits NOXO1-activation of Nox1. Thus, in transfected HEK293 cells, NOXO1 and NOXA1 activate Nox1 without the need for agonist activation, and this is mediated in part by binding of the NOXO1 PX domain to membrane lipids.  相似文献   

2.
In the plasma membrane fraction from Caco-2 human colon carcinoma cells, active Nox1 (NADPH oxidase 1) endogenously co-localizes with its regulatory components p22(phox), NOXO1, NOXA1 and Rac1. NADPH-specific superoxide generating activity was reduced by 80% in the presence of either a flavoenzyme inhibitor DPI (diphenyleneiodonium) or NADP(+). The plasma membranes from PMA-stimulated cells showed an increased amount of Rac1 (19.6 pmol/mg), as compared with the membranes from unstimulated Caco-2 cells (15.1 pmol/mg), but other components did not change before and after the stimulation by PMA. Spectrophotometric analysis found approx. 36 pmol of FAD and 43 pmol of haem per mg of membrane and the turnover of superoxide generation in a cell-free system consisting of the membrane and FAD was 10 mol/s per mol of haem. When the constitutively active form of Rac, Rac1(Q61L) or GTP-bound Rac1 was added exogenously to the membrane, O(2)(-)-producing activity was enhanced up to 1.5-fold above the basal level, but GDP-loaded Rac1 did not affect superoxide-generating kinetics. A fusion protein [NOXA1N-Rac1(Q61L)] between truncated NOXA1(1-211) and Rac1-(Q61L) exhibited a 6-fold increase of the basal Nox1 activity, but NOXO1N(1-292) [C-terminal truncated NOXO1(1-292)] alone showed little effect on the activity. The activated forms of Rac1 and NOXA1 are essentially involved in Nox1 activation and their interactions might be responsible for regulating the O(2)(-)-producing activity in Caco-2 cells.  相似文献   

3.
Rac1 has been implicated in the generation of reactive oxygen species (ROS) in several cell types, but the enzymatic origin of the ROS has not been proven. The present studies demonstrate that Nox1, a homolog of the phagocyte NADPH-oxidase component gp91(phox), is activated by Rac1. When Nox1 is co-expressed along with its regulatory subunits NOXO1 and NOXA1, significant ROS generation is seen. Herein, co-expression of constitutively active Rac1(G12V), but not wild-type Rac1, resulted in marked further stimulation of activity. Decreased Rac1 expression using small interfering RNA reduced Nox1-dependent ROS. CDC42(G12V) failed to increase activity, and small interfering RNA directed against CDC42 failed to decrease activity, pointing to specificity for Rac. TPR domain mutants of NOXA1 that interfere with Rac1 binding were ineffective in supporting Nox1-dependent ROS generation. Immunoprecipitation experiments demonstrated a complex containing Rac1(G12V), NOXO1, NOXA1, and Nox1. CDC42(G12V) could not substitute for Rac1(G12V) in such a complex. Nox1 formed a complex with Rac1(G12V) that was independent of NOXA1 and NOXO1, consistent with direct binding of Rac1(G12V) to Nox1. Rac1(G12V) interaction with NOXA1 was enhanced by Nox1 and NOXO1, suggesting cooperative binding. A model is presented comparing activation by regulatory subunits of Nox1 versus gp91(phox) (Nox2) in which Rac1 activation provides a major trigger that acutely activates Nox1-dependent ROS generation.  相似文献   

4.
gp91(phox) (Nox2), the catalytic subunit of the superoxide-generating respiratory burst oxidase, is regulated by subunits p47(phox) and p67(phox). Nox1, a homolog of gp91(phox), is regulated by NOXO1 and NOXA1, homologs of p47(phox) and p67(phox), respectively. For both Nox1 and gp91(phox), an organizer protein (NOXO1 or p47(phox)) cooperates with an activator protein (NOXA1 or p67(phox)) to regulate the catalytic subunit. Herein, we investigate the subunit regulation of Nox3 compared with that of other Nox enzymes. Nox3, like gp91(phox), was activated by p47(phox) plus p67(phox). Whereas gp91(phox) activity required the protein kinase C activator phorbol myristate acetate (PMA), Nox3 activity was already high without PMA, but was further stimulated approximately 30% by PMA. gp91(phox) was also activated by NOXO1/NOXA1 and required PMA for high activity. gp91(phox) regulation required an intact activation domain in the activator protein, as neither p67(phox)(V204A) nor NOXA1(V205A) were effective. In contrast, p67(phox)(V204A) was effective (along with p47(phox)) in activating Nox3. Unexpectedly, Nox3 was strongly activated by NOXO1 in the absence of NOXA1 or p67(phox). Nox3 activity was regulated by PMA only when p47(phox) but not NOXO1 was present, consistent with the phosphorylation-regulated autoinhibitory region in p47(phox) but not in NOXO1. Deletion of the autoinhibitory region from p47(phox) rendered this subunit highly active in the absence of PMA toward both gp91(phox) and Nox3, and high activity required an activator subunit. The unique regulation of Nox3 supports a model in which multiple interactions with regulatory subunits stabilize an active conformation of the catalytic subunit.  相似文献   

5.
Two novel proteins activate superoxide generation by the NADPH oxidase NOX1   总被引:16,自引:0,他引:16  
NOX1, an NADPH oxidase expressed predominantly in colon epithelium, shows a high degree of similarity to the phagocyte NADPH oxidase. However, superoxide generation by NOX1 has been difficult to demonstrate. Here we show that NOX1 generates superoxide when co-expressed with the p47(phox) and p67(phox) subunits of the phagocyte NADPH oxidase but not when expressed by itself. Since p47(phox) and p67(phox) are restricted mainly to myeloid cells, we searched for their homologues and identified two novel cDNAs. The mRNAs of both homologues were found predominantly in colon epithelium. Differences between the homologues and the phagocyte NADPH oxidase subunits included the lack of the autoinhibitory domain and the protein kinase C phosphorylation sites in the p47(phox) homologue as well as the absence of the first Src homology 3 domain and the presence of a hydrophobic stretch in the p67(phox) homologue. Co-expression of NOX1 with the two novel proteins led to stimulus-independent high level superoxide generation. Stimulus dependence of NOX1 was restored when p47(phox) was used to replace its homologue. In conclusion, NOX1 is a superoxide-generating enzyme that is activated by two novel proteins, which we propose to name NOXO1 (NOX organizer 1) and NOXA1 (NOX activator 1).  相似文献   

6.
ASAP1 (ADP ribosylation factor [ARF]- GTPase-activating protein [GAP] containing SH3, ANK repeats, and PH domain) is a phospholipid-dependent ARF-GAP that binds to and is phosphorylated by pp60(Src). Using affinity chromatography and yeast two-hybrid interaction screens, we identified ASAP1 as a major binding partner of protein tyrosine kinase focal adhesion kinase (FAK). Glutathione S-transferase pull-down and coimmunoprecipitation assays showed the binding of ASAP1 to FAK is mediated by an interaction between the C-terminal SH3 domain of ASAP1 with the second proline-rich motif in the C-terminal region of FAK. Transient overexpression of wild-type ASAP1 significantly retarded the spreading of REF52 cells plated on fibronectin. In contrast, overexpression of a truncated variant of ASAP1 that failed to bind FAK or a catalytically inactive variant of ASAP1 lacking GAP activity resulted in a less pronounced inhibition of cell spreading. Transient overexpression of wild-type ASAP1 prevented the efficient organization of paxillin and FAK in focal adhesions during cell spreading, while failing to significantly alter vinculin localization and organization. We conclude from these studies that modulation of ARF activity by ASAP1 is important for the regulation of focal adhesion assembly and/or organization by influencing the mechanisms responsible for the recruitment and organization of selected focal adhesion proteins such as paxillin and FAK.  相似文献   

7.
Consolidated evidence highlights the importance of redox signalling in poising the balance between self-renewal and differentiation in adult stem cells. The present study shows that human hematopoietic stem/progenitor cells (HSCs) constitutively generate low levels of hydrogen peroxide whose production is inhibited by DPI, apocynin, catalase, and LY294002 and scarcely stimulated by PMA. Moreover, it is shown that HSCs express at the mRNA and protein levels the catalytic subunits of NOX1, NOX2, and NOX4 isoforms of the NADPH oxidase family along with the complete battery of the regulatory subunits p22, p40, p47, p67, rac1, rac2, NOXO1, and NOXA1 as well as the splicing variant NOX2s and that the three NOX isoforms are largely co-expressed in the same HSC. These findings are interpreted in terms of a positive feed-back mechanism of NOXs activation enabling a fine tuning of the ROS level to be possibly used in redox-mediated signalling for growth and differentiation of HSCs.  相似文献   

8.
The integral membrane protein p22phox is an indispensable component of the superoxide-generating phagocyte NADPH oxidase, whose catalytic core is the membrane-associated gp91phox (also known as Nox2). p22phox associates with gp91phox and, through its proline-rich C terminus, provides a binding site for the tandem Src homology 3 domains of the activating subunit p47phox. Whereas p22phox is expressed ubiquitously, its participation in regulating the activity of other Nox enzymes is less clear. This study investigates the requirement of p22phox for Nox enzyme activity and explores the role of its proline-rich region (PRR) for regulating activity. Coexpression of specific Nox catalytic subunits (Nox1, Nox2, Nox3, Nox4, or Nox5) along with their corresponding regulatory subunits (NOXO1/NOXA1 for Nox1; p47phox/p67phox/Rac for Nox2; NOXO1 for Nox3; no subunits for Nox4 or Nox5) resulted in marked production of reactive oxygen. Small interfering RNAs decreased endogenous p22phox expression and inhibited reactive oxygen generation from Nox1, Nox2, Nox3, and Nox4 but not Nox5. Truncated forms of p22phox that disrupted the PRR-inhibited reactive oxygen generation from Nox1, Nox2, and Nox3 but not from Nox4 and Nox5. Similarly, p22phox (P156Q), a mutation that disrupts Src homology 3 binding by the PRR, potently inhibited reactive oxygen production from Nox1 and Nox2 but not from Nox4 and Nox5. Expression of p22phox (P156Q) inhibited NOXO1-stimulated Nox3 activity, but co-expression of NOXA1 overcame the inhibitory effect. The P157Q and P160Q mutations of p22phox showed selective inhibition of Nox2/p47phox/p67phox, and selectivity was specific for the organizing subunit (p47phox or NOXO1) rather than the Nox catalytic subunit. These studies stress the importance of p22phox for the function of Nox1, Nox2, Nox3, and Nox4, and emphasize the key role of the PRR for regulating Nox proteins whose activity is dependent upon p47phox or NOXO1.  相似文献   

9.
NOXO1β [NOXO1 (Nox organizer 1) β] is a cytosolic protein that, in conjunction with NOXA1 (Nox activator 1), regulates generation of reactive oxygen species by the NADPH oxidase 1 (Nox1) enzyme complex. NOXO1β is targeted to membranes through an N-terminal PX (phox homology) domain. We have used NMR spectroscopy to solve the structure of the NOXO1β PX domain and surface plasmon resonance (SPR) to assess phospholipid specificity. The solution structure of the NOXO1β PX domain shows greatest similarity to that of the phosphatidylinositol 3-kinase-C2α PX domain with regard to the positions and types of residues that are predicted to interact with phosphatidylinositol phosphate (PtdInsP) head groups. SPR experiments identify PtdIns(4,5)P(2) and PtdIns(3,4,5)P(3) as preferred targets of NOXO1β PX. These findings contrast with previous lipid overlay experiments showing strongest binding to monophosphorylated PtdInsP and phosphatidylserine. Our data suggest that localized membrane accumulation of PtdIns(4,5)P(2) or PtdIns(3,4,5)P(2) may serve to recruit NOXO1β and activate Nox1.  相似文献   

10.
The mechanism of angiotensin II (Ang II)-induced superoxide production was investigated with HEK293 or Chinese hamster ovary cells reconstituted with the angiotensin type 1 receptor (AT(1)R) and NADPH oxidase (either Nox1 or Nox2) along with a pair of adaptor subunits (either NOXO1 with NOXA1 or p47(phox) with p67(phox)). Ang II enhanced the activity of both Nox1 and Nox2 supported by either adaptor pair, with more effective activation of Nox1 in the presence of NOXO1 and NOXA1 and of Nox2 in the presence of p47(phox) and p67(phox). Expression of several AT(1)R mutants showed that interaction of the receptor with G proteins but not that with beta-arrestin or with other proteins (Jak2, phospholipase C-gamma1, SH2 domain-containing phosphatase 2) that bind to the COOH-terminal region of AT(1)R, was necessary for Ang II-induced superoxide production. The effects of constitutively active alpha subunits of G proteins and of various pharmacological agents implicated signaling by a pathway comprising AT(1)R, Galpha(q/11), phospholipase C-beta, and protein kinase C as largely, but not exclusively, responsible for Ang II-induced activation of Nox1 and Nox2 in the reconstituted cells. A contribution of Galpha(12/13), phospholipase D, and phosphatidyl-inositol 3-kinase to Ang II-induced superoxide generation was also suggested, whereas Src and the epidermal growth factor receptor did not appear to participate in this effect of Ang II. In reconstituted cells stimulated with Ang II, Nox2 exhibited a more sensitive response than Nox1 to the perturbation of protein kinase C, phosphatidylinositol 3-kinase, or the small GTPase Rac1.  相似文献   

11.
Kim MJ  Chang JS  Park SK  Hwang JI  Ryu SH  Suh PG 《Biochemistry》2000,39(29):8674-8682
A recent report that microinjection of the SH3 domain of PLC-gamma1 could induce DNA synthesis raised the functional importance of the SH3 domain of PLC-gamma1 in mitogenic signaling. In this report, we provide evidence that SOS1, a p21Ras-specific guanine nucleotide exchange factor, directly binds to the SH3 domain of PLC-gamma1, and that the SH3 domain of PLC-gamma1 is involved in SOS1-mediated p21Ras activation. SOS1 was coprecipitated with the GST-fused SH3 domain of PLC-gamma1 in vitro. The interaction between SOS1 and the PLC-gamma1 SH3 domain is mediated by direct physical interaction. The carboxyl-terminal proline-rich domain of SOS1 is involved in the interaction with the PLC-gamma1 SH3 domain. Moreover, PLC-gamma1 could be co-immunoprecipitated with SOS1 antibody in cell lysates. From transient expression studies, we could demonstrate that the SH3 domain of PLC-gamma1 is necessary for the association with SOS1 in vivo. Intriguingly, overexpression of the SH3 domain of PLC-gamma1, lipase-inactive PLC-gamma1, or wild-type PLC-gamma1 elevated p21Ras activity and ERK activity when compared with vector transfected cells. The PLC-gamma1 mutant lacking the SH3 domain could not activate p21Ras. p21Ras activities in cell lines overexpressing either PLC-gamma1 or the SH2-SH2-SH3 domain of PLC-gamma1 were elevated about 2-fold compared to vector transfected cells. This study is the first to demonstrate that the PLC-gamma1 SH3 domain enhances p21Ras activity, and that the SH3 domain of PLC-gamma1 may be involved in the SOS1-mediated signaling pathway.  相似文献   

12.
The production of reactive oxygen species (ROS) in mammalian cells is tightly regulated because of their potential to damage macromolecules, including DNA. To investigate possible links between high ROS levels, oxidative DNA damage, and genomic instability in mammalian cells, we established a novel model of chronic oxidative stress by coexpressing the NADPH oxidase human (h) NOX1 gene together with its cofactors NOXO1 and NOXA1. Transfectants of mismatch repair (MMR)-proficient HeLa cells or MMR-defective Msh2(-/-) mouse embryo fibroblasts overexpressing the hNOX1 complex displayed increased intracellular ROS levels. In one HeLa clone in which ROS were particularly elevated, reactive nitrogen species were also increased and nitrated proteins were identified with an anti-3-nitrotyrosine antibody. Overexpression of the hNOX1 complex increased the steady-state levels of DNA 8-oxo-7,8-dihydroguanine and caused a threefold increase in the HPRT mutation rate in HeLa cells. In contrast, additional oxidatively generated damage did not affect the constitutive mutator phenotype of the Msh2(-/-) fibroblasts. Because no significant changes in the expression of several DNA repair enzymes for oxidative DNA damage were identified, we suggest that chronic oxidative stress can saturate the cell's DNA repair capacity and cause significant genomic instability.  相似文献   

13.
In a screen for 3T3-F442A adipocyte proteins that bind SH2 domains, we isolated a cDNA encoding Fer, a nonreceptor protein-tyrosine kinase of the Fes/Fps family that contains a functional SH2 domain. A truncated splicing variant, iFer, was also cloned. iFer is devoid of both the tyrosine kinase domain and a functional SH2 domain but displays a unique 42-residue C terminus and retains the ability to form oligomers with Fer. Expression of both Fer and iFer proteins are strikingly increased upon differentiation of 3T3-L1 fibroblasts to adipocytes. Platelet-derived growth factor treatment of the cultured adipocytes caused rapid tyrosine phosphorylation of Fer and its recruitment to complexes containing platelet-derived growth factor receptor and the p85 regulatory subunit of phosphatidylinositol (PI) 3-kinase. Insulin treatment of 3T3-L1 adipocytes stimulated association of Fer with complexes containing tyrosine phosphorylated IRS-1 and PI 3-kinase but did not stimulate tyrosine phosphorylation of Fer. PI 3-kinase activity in anti-Fer immunoprecipitates was also acutely activated by insulin treatment of cultured adipocytes. These data demonstrate the presence of Fer tyrosine kinase in insulin signaling complexes, suggesting a role of Fer in insulin action.  相似文献   

14.
15.
Phospholipase-gamma1 (PLC-gamma1) prevents programmed cell death, for which the enzymatic activity has been implicated. However, the biological function of Src homology (SH) domains of PLC-gamma1 in promoting cell survival remains elusive. Here, we showed that deletion of the N-SH2 domain or both N-SH2 and C-SH2 domains, but not the SH3 domain, abolished the anti-apoptotic activity of PLC-gamma1. Surprisingly, removal of the whole SH domain inhibited apoptosis. The lipase-inactive PLC-gamma1 mutant (LIM) failed to suppress apoptosis. Moreover, the phospholipase activity in SH3- or whole SH domain-deleted cells was comparable to that of wild-type cells. By contrast, the enzymatic activity was substantially ablated in SH2 domain-deleted or LIM cells. A pharmacological inhibitor of PLC-gamma1 robustly diminished the anti-apoptotic action in wild-type, SH3- or whole SH domain-deleted cells, whereas pretreatment of SH2 domain-deleted or LIM cells with agents activating PKC and calcium mobilization markedly promoted cell survival. These results indicate that SH domains in PLC-gamma1 might mediate its anti-apoptotic action by regulating the enzymatic activity.  相似文献   

16.
Intramuscular immunization of mice with plasmids encoding two transdominant negative mutants of the HIV-1 Tat protein (Tat22 and Tat22/37) elicited a humoral response to wild-type Tat that is comparable to that induced by inoculation of wild-type tat DNA or Tat protein. The percentage of the responders and the Ab titers continued to increase after three additional DNA boosts and pretreatment with bupivacaine at the site of inoculation, without a significant difference (p > 0.05) among the three groups of mice immunized with mutant and wild-type tat genes. By utilizing synthetic peptides representing the amino acid sequence of Tat, one major B cell epitope was defined within the cysteine-rich domain of Tat. Anti-Tat IgG Abs directed against this epitope were found in mice immunized with all tat DNA constructs, whereas different Tat epitopes were detected in mice immunized with the Tat protein. Similarly, IgG2a was the predominant isotype in DNA-immunized mice, with both mutants and wild-type tat genes, as compared with protein immunization, which induced mostly IgG1 and IgG3. Sera from most immunized mice neutralized the effect of extracellular Tat in activating HIV-1 replication. A cellular response was also elicited as indicated by the proliferation of splenocytes when stimulated with wild-type Tat. These results indicate that the wild-type Tat Ag is recognized by Abs and T cells induced by DNA immunization with mutated tat genes, suggesting the possible use of these Tat transdominant mutants, lacking viral trans activation activity and capable of blocking wild-type Tat activity, in the development of an anti-HIV-1 vaccine.  相似文献   

17.
INI1/hSNF5 is an HIV-1 integrase (IN) binding protein specifically incorporated into virions. A truncated mutant of INI1 (S6, amino acids 183 to 294) harboring the minimal IN binding Rpt1 domain potently inhibits HIV-1 particle production in a transdominant manner. The inhibition requires interaction of S6 with IN within Gag-Pol. While INI1 is a nuclear protein and harbors a masked nuclear export signal (NES), the transdominant negative mutant S6 is cytoplasmic, due to the unmasking of NES. Here, we examined the effects of subcellular localization of S6 on HIV-1 inhibition and further investigated the stages of assembly that are affected. We found that targeting a nuclear localization signal-containing S6 variant [NLS-S6(Rpt1)] to the nucleoplasm (but not to the nucleolus) resulted in complete reversal of inhibition of particle production. Electron microscopy indicated that although no electron-dense particles at any stage of assembly were seen in cells expressing S6, virions were produced in cells expressing the rescue mutant NLS-S6(Rpt1) to wild-type levels. Immunofluorescence analysis revealed that p24 exhibited a diffuse pattern of localization within the cytoplasm in cells expressing S6 in contrast to accumulation along the membrane in controls. Pulse-chase analysis indicated that in S6-expressing cells, although Gag(Pr55(gag)) protein translation was unaffected, processing and release of p24 were defective. Together, these results indicate that expression of S6 in the cytoplasm interferes with trafficking of Gag-Pol/Gag to the membrane and causes a defective processing leading to inhibition of assembly at an early stage prior to particle formation and budding.  相似文献   

18.
A truncated form of the type 1 fibroblast growth factor receptor (FGFR1) lacking most of its cytoplasmic domain was tested for its ability to inhibit signal transduction by each of three different wild-type FGFRs (FGFR1, 2, and 3). When the truncated FGFR1 was expressed in Xenopus oocytes in excess of each wild-type FGFR, mobilization of intracellular calcium mediated by the wild-type FGFRs was completely blocked. The truncated FGFR did not inhibit signal transduction by the co-expressed platelet-derived growth factor beta-receptor. A form of truncated FGFR1 which lacked the first immunoglobulin-like domain also inhibited signal transduction by wild-type FGFRs. Truncated FGFR formed complexes with wild-type FGFR in the presence of basic FGF in intact cells. These observations were consistent with the hypothesis that the truncated FGFR interacted with wild-type FGFRs to form nonfunctional heterodimers, thus eliminating the signaling by the wild-type FGFRs. The observation that signaling by multiple types of FGFR can be blocked by a single type of truncated FGFR suggests that the different types of FGFR can interact with each other in ligand-mediated complexes. These findings provide a molecular basis for inhibiting the actions of FGFs in vivo.  相似文献   

19.
20.
SH2-Bbeta has been shown to bind via its SH2 (Src homology 2) domain to tyrosyl-phosphorylated JAK2 and strongly activate JAK2. In this study, we demonstrate the existence of an additional binding site(s) for JAK2 within the N-terminal region of SH2-Bbeta (amino acids 1 to 555) and the ability of this region of SH2-B to inhibit JAK2. Four lines of evidence support the existence of this additional binding site(s). In a glutathione S-transferase pull-down assay, wild-type SH2-Bbeta and SH2-Bbeta(R555E) with a defective SH2 domain bind to both tyrosyl-phosphorylated JAK2 from growth hormone (GH)-treated cells and non-tyrosyl-phosphorylated JAK2 from control cells, whereas the SH2 domain of SH2-Bbeta binds only to tyrosyl-phosphorylated JAK2 from GH-treated cells. Similarly, JAK2 is present in alphaSH2-B immunoprecipitates in the absence and presence of GH, with GH substantially increasing the coprecipitation of JAK2 with SH2-B. When coexpressed in COS cells, SH2-Bbeta coimmunoprecipitates not only wild-type, tyrosyl-phosphorylated JAK2 but also kinase-inactive, non-tyrosyl-phosphorylated JAK2(K882E), although to a lesser extent. DeltaC555 (amino acids 1 to 555 of SH2-Bbeta) that lacks most of the SH2 domain binds similarly to wild-type JAK2 and kinase-inactive JAK2(K882E). Experiments using a series of N- and C-terminally truncated SH2-Bbeta constructs indicate that the pleckstrin homology (PH) domain (amino acids 269 to 410) and amino acids 410 to 555 are necessary for maximal binding of SH2-Bbeta to inactive JAK2, but neither region alone is sufficient for maximal binding. The SH2 domain of SH2-Bbeta is necessary and sufficient for the stimulatory effect of SH2-Bbeta on JAK2 and JAK2-mediated tyrosyl phosphorylation of Stat5B. In contrast, DeltaC555 lacking the SH2 domain, and to a lesser extent the PH domain alone, inhibits JAK2. DeltaC555 also blocks JAK2-mediated tyrosyl phosphorylation of Stat5B in COS cells and GH-stimulated nuclear accumulation of Stat5B in 3T3-F442A cells. These data indicate that in addition to the SH2 domain, SH2-Bbeta has one or more lower-affinity binding sites for JAK2 within amino acids 269 to 555. The interaction via this site(s) in SH2-B with inactive JAK2 seems likely to increase the local concentration of SH2-Bbeta around JAK2, thereby facilitating binding of the SH2 domain to ligand-activated JAK2. This would result in a more rapid and robust cellular response to hormones and cytokines that activate JAK2. This interaction between inactive JAK2 and SH2-B may also help prevent abnormal activation of JAK2.  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