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.
High-risk types of human papillomavirus (HPV) are considered the major causative agents of cervical carcinoma. The transforming ability of HPV resides in the E6 and E7 oncogenes, yet the pathway to transformation is not well understood. Cells expressing the oncogene E7 from high-risk HPVs have a high incidence of polyploidy, which has been shown to occur as an early event in cervical carcinogenesis and predisposes the cells to aneuploidy. The mechanism through which E7 contributes to polyploidy is not known. It has been hypothesized that E7 induces polyploidy in response to mitotic stress by abrogating the mitotic spindle assembly checkpoint. It was also proposed that E7 may stimulate rereplication to induce polyploidy. We have tested these hypotheses by using human epithelial cells in which E7 expression induces a significant amount of polyploidy. We find that E7-expressing cells undergo normal mitoses with an intact spindle assembly checkpoint and that they are able to complete cytokinesis. Our results also exclude DNA rereplication as a major mechanism of polyploidization in E7-expressing cells upon microtubule disruption. Instead, we have shown that while normal cells arrest at the postmitotic checkpoint after adaptation to the spindle assembly checkpoint, E7-expressing cells replicate their DNA and propagate as polyploid cells. Thus, abrogation of the postmitotic checkpoint leads to polyploidy formation in E7-expressing human epithelial cells. Our results suggest that downregulation of pRb is important for E7 to induce polyploidy and abrogation of the postmitotic checkpoint.An important hallmark of human cancers is aneuploidy, the state in which a cell has extra or missing chromosomes (12, 25). Polyploidy is the state in which cells have more than two equal sets of chromosomes and is thought to be an early event in multistep carcinogenesis that can lead to aneuploidy (1, 24), as exemplified in Barrett''s esophagus (11). Polyploidy has recently been shown to occur as an early event in cervical carcinogenesis and to predispose the cells to aneuploidy (26). Other recent studies have shown that tetraploid but not diploid mouse or human cells induce tumor formation in mice (3, 9). These studies highlight the potential importance of polyploidy in carcinogenesis.The cellular mechanisms responsible for this polyploidy formation are as of yet undetermined, but several models have been proposed. First, abrogation of the spindle assembly checkpoint followed by cleavage failure may lead to polyploidy formation (36, 40). A second proposed model is rereplication, a process of multiple rounds of DNA replication without an intervening mitosis. Third, cells that adapt to the mitotic spindle checkpoint halt in a G1-like state with 4C DNA content. Abrogation of this postmitotic checkpoint allows the cells to replicate their 4C DNA content, leading to polyploidy formation. This has been shown in cells that express the human papillomavirus type 16 (HPV-16) E6 oncogene that degrades p53 (21). Finally, cleavage failure, which yields binucleate cells with 4C DNA content, is also a potential mechanism for polyploidy formation (31).The postmitotic checkpoint becomes activated when cells with an intact spindle assembly checkpoint become arrested during mitosis for a prolonged period of time and eventually adapt to the checkpoint, exit mitosis without cleavage, and progress into a G1-like state with 4C DNA content (19, 22). The cells are prevented from continuing through the cell cycle and replicating their DNA by a proposed p53- and pRb-dependent postmitotic checkpoint (18, 19).High-risk types of HPV (of which HPV-16 is the most prevalent) are commonly associated with lesions that can progress to cervical carcinoma, which is one of the leading causes of cancer death in women worldwide (42). The transforming properties of high-risk HPVs primarily reside in the E6 and E7 oncogenes (reviewed in reference 7). The ability of high-risk HPV E6 and E7 proteins to promote the degradation of p53 and pRb, respectively, has been suggested as a mechanism by which HPV induces cellular transformation (6, 30). Expression of the high-risk HPV E6 and E7 oncogenes in human keratinocytes leads to polyploidy, which is enhanced by DNA damage and by activation of the spindle checkpoint through microtubule disruption (15, 27, 37, 38).Previously, it was thought but not directly shown that high-risk E6 and E7 induce polyploidy in response to microtubule disruption by abrogating the spindle checkpoint and that degradation of the tumor suppressor p53 by E6 is the mechanism by which E6 accomplishes this polyploidy formation (27, 37, 38). Others have proposed that E7 may play a role in stimulating DNA rereplication that occurs prior to mitosis initiation and polyploidy formation (20). Our recent studies demonstrate that E6 does not affect the mitotic spindle checkpoint (21). Instead, E6 abrogates the postmitotic checkpoint to induce polyploidy after microtubule disruption. Interestingly, E6 mutant proteins defective in inducing p53 degradation also induce polyploidy (21). The mechanism by which HPV E7 induces polyploidy remains to be determined. In this study, we investigate these possible mechanisms through which HPV-16 E7 induces polyploidy formation.  相似文献   

2.
3.
Mitotic centromere-associated kinesin (MCAK) plays an essential role in spindle formation and in correction of improper microtubule-kinetochore attachments. The localization and activity of MCAK at the centromere/kinetochore are controlled by Aurora B kinase. However, MCAK is also abundant in the cytosol and at centrosomes during mitosis, and its regulatory mechanism at these sites is unknown. We show here that cyclin-dependent kinase 1 (Cdk1) phosphorylates T537 in the core domain of MCAK and attenuates its microtubule-destabilizing activity in vitro and in vivo. Phosphorylation of MCAK by Cdk1 promotes the release of MCAK from centrosomes and is required for proper spindle formation. Interfering with the regulation of MCAK by Cdk1 causes dramatic defects in spindle formation and in chromosome positioning. This is the first study demonstrating that Cdk1 regulates the localization and activity of MCAK in mitosis by directly phosphorylating the catalytic core domain of MCAK.Chromosomes are properly attached to the mitotic spindles, and chromosome movement is tightly linked to the structure and dynamics of spindle microtubules during mitosis. Important regulators of microtubule dynamics are the kinesin-13 proteins (37). This kinesin superfamily is defined by the localization of the conserved kinesin core motor domain in the middle of the polypeptide (19). Kinesin-13 proteins induce microtubule depolymerization by disassembling tubulin subunits from the polymer end (6). Among them, mitotic centromere-associated kinesin (MCAK) is the best-characterized member of the family. It depolymerizes microtubules in vitro and in vivo, regulates microtubule dynamics, and has been implicated in correcting misaligned chromosomes (12, 14, 16, 24). In agreement with these observations, both overexpression and inhibition of MCAK result in a disruption of microtubule dynamics, leading further to improper spindle assembly and errors in chromosome alignment and segregation (7, 11, 15, 22, 33). The importance of MCAK in ensuring the faithful segregation of chromosomes is consistent with the observation that MCAK is highly expressed in several types of cancer and thus is likely to be involved in causing aneuploidy (25, 32).While MCAK is found both in the cytoplasm and at the centromeres throughout the cell cycle, it is highly enriched on centrosomes, the centromeres/kinetochores, and the spindle midzone during mitosis (18, 21, 36, 38). In accordance with its localizations, MCAK affects many aspects throughout mitosis, from spindle assembly and maintenance (3, 10, 36) to chromosome positioning and segregation (14, 21, 35). Thus, the precise control of the localization and activity of MCAK is crucial for maintaining genetic integrity during mitosis. Regulation of MCAK on the centromeres/kinetochores by Aurora B kinase in mitosis has been intensively investigated (1, 28, 29, 43). The data reveal that MCAK is phosphorylated on several serine/threonine residues by Aurora B, which inhibits the microtubule-destabilizing activity of MCAK and regulates its localization on chromosome arms/centromeres/kinetochores during mitosis (1, 18, 28). Moreover, in concert with Aurora B, ICIS (inner centromere KinI stimulator), a protein targeting the inner centromeres in an MCAK-dependent manner, may regulate MCAK at the inner centromeres and prevent kinetochore-microtubule attachment errors in mitosis by stimulating the activity of MCAK (27). Interestingly, hSgo2, a recently discovered inner centromere protein essential for centromere cohesion, has been reported to be important in localizing MCAK to the centromere and in spatially regulating its mitotic activity (13). These data highlight that the activity and localization of MCAK on the centromeres/kinetochores during mitosis are tightly controlled by Aurora B and its cofactors. Remarkably, MCAK concentrates at spindle poles from prophase to telophase during mitosis (18); however, only a few studies have been done to deal with that issue. Aurora A-depleted prometaphase cells delocalize MCAK from spindle poles but accumulate the microtubule-stabilizing protein ch-TOG at poles (5), implying that Aurora A might influence the centrosomal localization of MCAK in mitosis. Aurora A is also found to be important for focusing microtubules at aster centers and for facilitating the transition from asters to bipolar spindles in Xenopus egg extracts (42). In addition, it has been revealed that Ca2+/calmodulin-dependent protein kinase II gamma (CaMKII gamma) suppresses MCAK''s activity, which is essential for bipolar spindle formation in mitosis (11). More work is required to gain insight into the regulatory mechanisms of MCAK at spindle poles during mitosis.Deregulated cyclin-dependent kinases (Cdks) are very often linked to genomic and chromosomal instability (20). Cyclin B1, the regulatory subunit of Cdk1, is localized to unattached kinetochores and contributes to efficient microtubule attachment and proper chromosome alignment (2, 4). We observed that knockdown of cyclin B1 induces defects in chromosome alignment and mitotic spindle formation (N.-N. Kreis, M. Sanhaji, A. Krämer, K. Sommor, F. Rödel, K. Strebhardt, and J. Yuan, submitted for publication). Yet, how Cdk1/cyclin B1 carries out these functions is not very well understood. In this context, it is extremely interesting to investigate the relationship between the essential mitotic kinase Cdk1 and the microtubule depolymerase MCAK in human cells.  相似文献   

4.
Despite detailed knowledge of the components of the spindle assembly checkpoint, a molecular explanation of how cells die after prolonged spindle checkpoint activation, and thus how microtubule inhibitors and other antimitotic drugs ultimately elicit their lethal effects, has yet to emerge. Mitotically arrested cells typically display extensive phosphorylation of two key antiapoptotic proteins, Bcl-xL and Bcl-2, and evidence suggests that phosphorylation disables their antiapoptotic activity. However, the responsible kinase has remained elusive. In this report, evidence is presented that cyclin-dependent kinase 1 (CDK1)/cyclin B catalyzes mitotic-arrest-induced Bcl-xL/Bcl-2 phosphorylation. Furthermore, we show that CDK1 transiently and incompletely phosphorylates these proteins during normal mitosis. When mitosis is prolonged in the absence of microtubule inhibition, Bcl-xL and Bcl-2 become highly phosphorylated. Transient overexpression of nondegradable cyclin B1 caused apoptotic death, which was blocked by a phosphodefective Bcl-xL mutant but not by a phosphomimetic Bcl-xL mutant, confirming Bcl-xL as a key target of proapoptotic CDK1 signaling. These findings suggest a model whereby a switch in the duration of CDK1 activation, from transient during mitosis to sustained during mitotic arrest, dramatically increases the extent of Bcl-xL/Bcl-2 phosphorylation, resulting in inactivation of their antiapoptotic function. Thus, phosphorylation of antiapoptotic Bcl-2 proteins acts as a sensor for CDK1 signal duration and as a functional link coupling mitotic arrest to apoptosis.The cell division cycle is controlled by checkpoints, which ensure the fidelity of chromosome replication and segregation, as well as orderly progression through the cell cycle. If these critical events cannot be completed as scheduled, damaged cells, which might otherwise pose a threat to the organism as precancerous cells, are eliminated (16). The mitotic checkpoint, for example, produces a “prevent anaphase” signal until all the chromosomes are properly attached to kinetochores (22). Microtubule inhibitors (MTIs) and other antimitotic agents prolong the activation of this checkpoint, causing mitotic arrest, which culminates in cell death generally via intrinsic apoptosis, providing a rationale for the use of these agents as antitumor agents (20, 31). Intrinsic or mitochondrial apoptosis is regulated by the Bcl-2 family of proteins, which exhibit either pro- or antiapoptotic properties (17, 37). The BH3-only proapoptotic members act as essential initiators of intrinsic apoptosis, whereas the multidomain proapoptotic members, Bax and Bak, act as essential mediators of mitochondrial membrane permeability. Antiapoptotic Bcl-2 family members, including Bcl-xL, Bcl-2, and Mcl-1, oppose apoptosis by binding to the proapoptotic members and neutralizing their activity.The molecular mechanisms leading to cell death in response to spindle checkpoint activation have yet to be established. Indeed, how the spindle checkpoint couples to pathways regulating cell survival and death still represents an unresolved issue in cell biology (26, 35). Nonetheless, it seems reasonable to hypothesize that signals generated in response to prolonged mitotic arrest are eventually transduced to the apoptotic machinery. In this regard, it is striking that MTIs consistently induce the phosphorylation of two key antiapoptotic proteins, Bcl-2 and Bcl-xL, whereas other apoptotic stimuli fail to do so (9, 13, 25). The results of studies with phosphodefective mutants of Bcl-2 and Bcl-xL indicate that phosphorylation antagonizes their antiapoptotic function (2, 33, 36), but the precise mechanism(s) has yet to be fully clarified.The identity of the kinase responsible for the extensive phosphorylation of Bcl-xL and Bcl-2 that occurs in response to sustained spindle checkpoint activation is unresolved. Identification of this kinase is considered to be of critical importance, since it will provide insight into the molecular links between mitotic arrest and cell death, as well as the molecular mechanism of action of antimitotic drugs. Several candidates have been proposed, including Raf-1 (3), Jun N-terminal protein kinase (JNK) (2, 11, 36), protein kinase A (PKA) (32), cyclin-dependent kinase 1 (CDK1) (24), and mammalian target of rapamycin (mTOR) (4). In general, however, conclusions have been correlative or have been based on the use of kinase inhibitors tested under conditions that precluded mitotic arrest and thus indirectly blocked the effects of MTIs. Thus, strong experimental evidence supporting identification is lacking.Here we present evidence that the CDK1/cyclin B kinase complex is responsible for mitotic arrest-induced Bcl-xL/Bcl-2 phosphorylation. Furthermore, we show that CDK1 transiently and incompletely phosphorylates these proteins during normal mitosis. The findings suggest a model whereby a switch in the duration of CDK1 activation, from transient during mitosis to sustained during mitotic arrest, dramatically increases the extent of Bcl-xL/Bcl-2 phosphorylation, resulting in inactivation of the antiapoptotic function of Bcl-xL/Bcl-2. Thus, CDK1-mediated phosphorylation of antiapoptotic Bcl-2 proteins acts as a key link coupling mitotic arrest to apoptosis.  相似文献   

5.
Mrc1 plays a role in mediating the DNA replication checkpoint. We surveyed replication elongation proteins that interact directly with Mrc1 and identified a replicative helicase, Mcm6, as a specific Mrc1-binding protein. The central portion of Mrc1, containing a conserved coiled-coil region, was found to be essential for interaction with the 168-amino-acid C-terminal region of Mcm6, and introduction of two amino acid substitutions in this C-terminal region abolished the interaction with Mrc1 in vivo. An mcm6 mutant bearing these substitutions showed a severe defect in DNA replication checkpoint activation in response to stress caused by methyl methanesulfonate. Interestingly, the mutant did not show any defect in DNA replication checkpoint activation in response to hydroxyurea treatment. The phenotype of the mcm6 mutant was suppressed when the mutant protein was physically fused with Mrc1. These results strongly suggest for the first time that an Mcm helicase acts as a checkpoint sensor for methyl methanesulfonate-induced DNA damage through direct binding to the replication checkpoint mediator Mrc1.Progression of the DNA replication machinery along chromosomes is a complex process. Replication forks pause occasionally when they encounter genomic regions that are difficult to replicate, such as highly transcribed regions, tRNA genes, and regions with specialized chromatin structure, like centromeric and heterochromatic regions (17). Replication forks also stall when treated with chemicals like methyl methanesulfonate (MMS), which causes DNA damage, or hydroxyurea (HU), which limits the cellular concentration of the deoxynucleoside triphosphate pool (17). Because de novo assembly and programming of the replisome do not occur after the onset of S phase (18), DNA replication forks must be protected from replicative stresses. The DNA replication checkpoint constitutes a surveillance mechanism for S-phase progression that safeguards replication forks from various replicative stresses (22, 38, 40), and malfunction of this checkpoint leads to chromosome instability and cancer development in higher organisms (4, 9).The Saccharomyces cerevisiae DNA replication checkpoint mediator Mrc1 is functionally conserved and is involved directly in DNA replication as a component of the replisome (1, 8, 16, 19, 29, 30). Mrc1, together with Tof1 and Csm3, is required for forming a replication pausing complex when the fork is exposed to replicative stress by HU (16). The pausing complex subsequently triggers events leading to DNA replication checkpoint activation and hence stable replicative arrest. A sensor kinase complex, Mec1-Ddc2 (ATR-ATRIP homolog of higher eukaryotes), is then recruited to the complex (14, 16). Mec1-Ddc2-mediated phosphorylation of Mrc1 activates the pausing complex, and phosphorylated Mrc1 likely recruits Rad53 (a putative homolog of CHK2 of higher eukaryotes), which is then activated via phosphorylation by Mec1-Ddc2 (1, 16, 20, 30). Activated Rad53 subsequently elicits a stress responses, i.e., stabilization of replication forks, induction of repair genes, and suppression of late-firing origins (24). It remains unclear, however, whether DNA replication checkpoint activation is induced in response to DNA damage by MMS, a reagent commonly used to study the DNA replication stress response. Several lines of evidence have suggested that MMS-induced damage is also sensed directly by the replication machinery (38, 40).Although biochemical and genetic interaction data have placed Mrc1 at the center of the replication checkpoint signal transduction cascade, its molecular function remains largely unknown. The proteins Mrc1, Tof1, and Csm3 associate with the Mcm complex (8, 27), a heterohexameric DNA helicase consisting of Mcm2 to Mcm7 proteins which unwinds the parental DNA duplex to allow replisome progression (3, 12, 18, 31, 32, 35). The Mcm complex associates with a specific set of regulatory proteins at forks to form replisome progression complexes (8). In addition to Mcm, Tof1, Csm3, and Mrc1, replisome progression complexes include factors such as Cdc45 and the GINS complex that are also required for fork progression (13, 26, 31, 32, 39). Claspin, a putative Xenopus laevis homolog of Mrc1, is also reported to associate with Cdc45, DNA polymerase ɛ (Polɛ), replication protein A, and two of the replication factor C complexes in aphidicolin-treated Xenopus egg extracts (19). Recently, Mrc1 was reported to interact directly with Polɛ (23).The aim of this study was to provide mechanistic insight into Mrc1 function in the DNA replication checkpoint. For this purpose, it was essential to identify, among all the essential proteins in the replication machinery, a specific protein that interacts with Mrc1 and to examine the role of this interaction in the DNA replication checkpoint. We found that Mrc1 interacts with Mcm6 directly and specifically. When the interaction between Mrc1 and Mcm6 was impaired, cells no longer activated the DNA replication checkpoint in response to MMS-induced replicative stress. Interestingly and unexpectedly, this interaction was not required for DNA replication checkpoint activation in response to HU-induced replicative stress. Our results provide the first mechanistic evidence that cells use separate mechanisms to transmit replicative stresses caused by MMS and HU for DNA replication checkpoint activation.  相似文献   

6.
7.
8.
9.
Cells expressing human papillomavirus type 16 (HPV-16) E6 and E7 proteins exhibit deregulation of G2/M genes, allowing bypass of DNA damage arrest signals. Normally, cells with DNA damage that override the G2 damage checkpoint would precociously enter mitosis and ultimately face mitotic catastrophe and apoptotic cell death. However, E6/E7-expressing cells (E6/E7 cells) have the ability to enter and exit mitosis in the presence of DNA damage and continue with the next round of the cell cycle. Little is known about the mechanism that allows these cells to gain entry into and exit from mitosis. Here, we show that in the presence of DNA damage, E6/E7 cells have elevated levels of cyclin B, which would allow entry into mitosis. Also, as required for exit from mitosis, cyclin B is degraded in these cells, permitting initiation of the next round of DNA synthesis and cell cycle progression. Proteasomal degradation of cyclin B by anaphase-promoting complex/cyclosome (APC/C) is, in part, due to elevated levels of the E2-conjugating enzyme, Ubch10, and the substrate recognition protein, Cdc20, of APC/C. Also, in E6/E7 cells with DNA damage, while Cdc20 is complexed with BubR1, indicating an active checkpoint, it is also present in complexes free of BubR1, presumably allowing APC/C activity and slippage through the checkpoint.Failure to activate cell cycle checkpoints in the presence of any DNA damage leads to genomic instability, polyploidy, and subsequently, aneuploidy, which is a hallmark of many cancers (26). Human papillomaviruses (HPVs) which cause various epithelial cancers, produce two proteins, E6 and E7, whose expression allows bypass or overriding of normal DNA damage and spindle checkpoint signals, primarily through inactivation of p53 and retinoblastoma family members, respectively (11, 16, 17). Our laboratory and others have previously shown that bypass of these arrest signals due to the presence of the viral genes gives rise to a significant population of cells that are polyploid (13, 16, 24, 32). Polyploid and aneuploid cells predominantly arise due to defects in the spindle assembly checkpoint (SAC) during mitosis. While we have some understanding of the mechanisms that lead to bypass of DNA damage arrest signals at the G2/M stage of the cell cycle, it is not clear how the E6/E7-expressing cells with DNA damage and abnormal chromosomes are allowed to (i) to enter into mitosis and (ii) exit out of mitosis to initiate the next round of replication. Progression through mitosis is regulated by the ubiquitin-dependent degradation machinery, consisting of the anaphase-promoting complex/cyclosome (APC/C), a multisubunit ubiquitin ligase. The activity of APC/C is dependent on the substrate-specifying proteins Cdc20 in metaphase and Cdh1 in telophase (25, 37). In normal cells, spindle checkpoint proteins Mad2 and BubR1 serve to inhibit APC/C until all the chromosomes are aligned correctly on the mitotic spindle by binding Cdc20 and preventing it from activating APC/C (5, 21, 31). In the event of DNA damage and/or unattached kinetochores, the SAC will arrest cells before exit from mitosis by inhibiting activation of APC/C. As a consequence of APC/C inhibition, cyclin B is not degraded, thus preventing cells from mitotic exit (6). Work by Chen''s group (11) has shown that E6- and E7-expressing cells (also referred to here as E6/E7 cells) adapt to an active SAC and are capable of mitotic slippage. So, what is the mechanism that underlies mitotic slippage in E6/E7 cells and allows them to enter the next round of cell cycle? Recent work by van Ree et al. (34) has shown that overexpression of E2 ubiquitin-conjugating enzyme Ubch10 leads to uncontrolled APC/C activity and degradation of cyclin B even in the presence of an active mitotic checkpoint, leading to mitotic slippage. In this report, we show that primary human foreskin keratinocytes (HFKs) expressing E6/E7 have high levels of cyclin B, which allows entry into mitosis in the presence of DNA damage. We show that these cells successfully exit mitosis by, in part, indirect activation of APC/C through upregulation of the E2-conjugating protein, Ubch10, and the substrate-specific component of APC/C, Cdc20, leading to the required degradation of cyclin B. In addition, Cdc20 is detected in different complexes; one includes the protein BubR1, indicating an active checkpoint, while other complexes are free of BubR1 and are thus free to activate APC/C. Upregulation of cyclin B and Ubch10 as well as Cdc20 is primarily through E6 and its ability to target p53 degradation, although inhibition of the pRb family members by E7 may also play a part.  相似文献   

10.
The Asf1 and Rad6 pathways have been implicated in a number of common processes such as suppression of gross chromosomal rearrangements (GCRs), DNA repair, modification of chromatin, and proper checkpoint functions. We examined the relationship between Asf1 and different gene products implicated in postreplication repair (PRR) pathways in the suppression of GCRs, checkpoint function, sensitivity to hydroxyurea (HU) and methyl methanesulfonate (MMS), and ubiquitination of proliferating cell nuclear antigen (PCNA). We found that defects in Rad6 PRR pathway and Siz1/Srs2 homologous recombination suppression (HRS) pathway genes suppressed the increased GCR rates seen in asf1 mutants, which was independent of translesion bypass polymerases but showed an increased dependency on Dun1. Combining an asf1 deletion with different PRR mutations resulted in a synergistic increase in sensitivity to chronic HU and MMS treatment; however, these double mutants were not checkpoint defective, since they were capable of recovering from acute treatment with HU. Interestingly, we found that Asf1 and Rad6 cooperate in ubiquitination of PCNA, indicating that Rad6 and Asf1 function in parallel pathways that ubiquitinate PCNA. Our results show that ASF1 probably contributes to the maintenance of genome stability through multiple mechanisms, some of which involve the PRR and HRS pathways.DNA replication must be highly coordinated with chromatin assembly and cell division for correct propagation of genetic information and cell survival. Errors arising during DNA replication are corrected through the functions of numerous pathways including checkpoints and a diversity of DNA repair mechanisms (32, 33, 35). However, in the absence of these critical cellular responses, replication errors can lead to the accumulation of mutations and gross chromosomal rearrangements (GCRs) as well as chromosome loss, a condition generally termed genomic instability (33). Genome instability is a hallmark of many cancers as well as other human diseases (24). There are many mechanisms by which GCRs can arise, and over the last few years numerous genes and pathways have been implicated in playing a role in the suppression of GCRs in Saccharomyces cerevisiae and in some cases in the etiology of cancer (27, 28, 33, 39-47, 51, 53, 56, 58, 60), including S. cerevisiae ASF1, which encodes the main subunit of the replication coupling assembly factor (37, 62).Asf1 is involved in the deposition of histones H3 and H4 onto newly synthesized DNA during DNA replication and repair (62), and correspondingly, asf1 mutants are sensitive to chronic treatment with DNA-damaging agents (2, 30, 62). However, asf1 mutants do not appear to be repair defective and can recover from acute treatment with at least some DNA-damaging agents (2, 8, 30, 31, 54), properties similar to those described for rad9 mutants (68). In the absence of Asf1, both the DNA damage and replication checkpoints become activated during normal cell growth, and in the absence of checkpoint execution, there is a further increase in checkpoint activation in asf1 mutants (30, 46, 54). It has been suggested that asf1 mutants are defective for checkpoint shutoff and that this might account for the increased steady-state levels of checkpoint activation seen in asf1 mutants (8); however, another study has shown that asf1 mutants are not defective for checkpoint shutoff and that in fact Asf1 and the chromatin assembly factor I (CAF-I) complex act redundantly or cooperate in checkpoint shutoff (31). Furthermore, Asf1 might be involved in proper activation of the Rad53 checkpoint protein, as Asf1 physically interacts with Rad53 and this interaction is abrogated in response to exogenous DNA damage (15, 26); however, the physiological relevance of this interaction is unclear. Asf1 is also required for K56 acetylation of histone H3 by Rtt109, and both rtt109 mutants and histone H3 variants that cannot be acetylated (38) share many of the properties of asf1 mutants, suggesting that at least some of the requirement for Asf1 in response to DNA damage is mediated through Rtt109 (11, 14, 22, 61). Subsequent studies of checkpoint activation in asf1 mutants have led to the hypothesis that replication coupling assembly factor defects result in destabilization of replication forks which are then recognized by the replication checkpoint and stabilized, suggesting that the destabilized replication forks account for both the increased GCRs and increased checkpoint activation seen in asf1 mutants (30). This hypothesis is supported by other recent studies implicating Asf1 in the processing of stalled replication forks (16, 57). This role appears to be independent of CAF-I, which can cooperate with Asf1 in chromatin assembly (63). Asf1 has also been shown to function in disassembly of chromatin, suggesting other possibilities for the mechanism of action of Asf1 at the replication fork (1, 2, 34). Thus, while Asf1 is thought to be involved in progression of the replication fork, both the mechanism of action and the factors that cooperate with Asf1 in this process remain obscure.Stalled replication forks, particularly those that stall at sites of DNA damage, can be processed by homologous recombination (HR) (6) or by a mechanism known as postreplication repair (PRR) (reviewed in reference 67). There are two PRR pathways, an error-prone pathway involving translesion synthesis (TLS) by lower-fidelity polymerases and an error-free pathway thought to involve template switching (TS) (67). In S. cerevisiae, the PRR pathways are under the control of the RAD6 epistasis group (64). The error-prone pathway depends on monoubiquitination of proliferating cell nuclear antigen (PCNA) on K164 by Rad6 (an E2 ubiquitin-conjugating enzyme) by Rad18 (E3 ubiquitin ligase) (23). This results in replacement of the replicative DNA polymerase with nonessential TLS DNA polymerases, such as REV3/REV7-encoded DNA polymerase ζ (polζ) and RAD30-encoded DNA polη, which can bypass different types of replication-blocking damage (67). The error-free pathway is controlled by Rad5 (E3) and a complex consisting of Ubc13 and Mms2 (E2 and E2 variant, respectively), which add a K63-linked polyubiquitin chain to monoubiquitinated PCNA, leading to TS to the undamaged nascent sister chromatid (4, 25, 65). Furthermore, in addition to modification with ubiquitin, K164 of PCNA can also be sumoylated by Siz1, resulting in subsequent recruitment of the Srs2 helicase and inhibition of deleterious Rad51-dependent recombination events (50, 52, 55), although it is currently unclear if these are competing PCNA modifications or if both can exist on different subunits in the same PCNA trimer. A separate branch of the Rad6 pathway involving the E3 ligase Bre1 monoubiquitinates the histone H2B (29, 69) as well as Swd2 (66), which stimulates Set1-dependent methylation of K4 and Dot1-dependent methylation of K79 of histone H3 (48, 49, 66). Subsequently, K79-methylated H3 recruits Rad9 and activates the Rad53 checkpoint (19, 70). Activation of Rad53 is also bolstered by Rad6-Rad18-dependent ubiquitination of Rad17, which is part of the 9-1-1 complex that functions upstream in the checkpoint pathway (17). Finally, Rad6 complexes with the E3 Ubr1, which mediates protein degradation by the N-end rule pathway (13).Due to the role of the PRR pathways at stalled replication forks and a recent study implicating the Rad6 pathway in the suppression of GCRs (39), we examined the relationship between these ubiquitination and sumoylation pathways and the Asf1 pathway in order to gain additional insights into the function of Asf1 during DNA replication and repair. Our findings suggest that Asf1 has multiple functions that prevent replication damage or act in the cellular responses to replication damage and that these functions are modified by and interact with the PRR pathways. The TLS PRR pathway does not appear to be involved, and both a Dun1-dependent replication checkpoint and HR are important for preventing the deleterious effects of PRR and Asf1 pathway defects. We hypothesize that this newly observed cooperation between Asf1 and the PRR pathways may be required for resolving stalled replication forks, leading to suppression of GCRs and successful DNA replication.  相似文献   

11.
12.
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   

13.
To prevent aneuploidy, cells require a mitotic surveillance mechanism, the spindle assembly checkpoint (SAC). The SAC prevents metaphase/anaphase transition by blocking the ubiquitylation and destruction of cyclin B and securin via the Cdc20-activated anaphase-promoting complex or cyclosome (APC/C)-mediated proteolysis pathway. This checkpoint involves the kinetochore proteins Mad2, BubR1, and Cdc20. Mad2 and BubR1 are inhibitors of the APC/C, but Cdc20 is an activator. Exactly how the SAC regulates Cdc20 via unattached kinetochores remains unclear; in vertebrates, most current models suggest that kinetochore-bound Mad2 is required for initial binding to Cdc20 to form a stable complex that includes BubR1. Here, we show that the Mad2 kinetochore dimerization recruitment mechanism is conserved and that the recruitment of Cdc20 to kinetochores in Drosophila requires BubR1 but not Mad2. BubR1 and Mad2 can bind to Cdc20 independently, and the interactions are enhanced after cells are arrested at mitosis by the depletion of Cdc27 using RNA interference (RNAi) in S2 cells or by MG132 treatment in syncytial embryos. These findings offer an explanation of why BubR1 is more important than Mad2 for SAC function in flies. These findings could lead to a better understanding of vertebrate SAC mechanisms.The spindle assembly checkpoint (SAC) is a mitotic surveillance mechanism that negatively regulates the activation of the anaphase-promoting complex or cyclosome (APC/C)-mediated proteolysis pathway to prevent the destruction of two key substrates, cyclin B and securin, thereby inhibiting the metaphase-to-anaphase transition until bipolar attachment of all chromosomes has been achieved (35). A number of conserved kinetochore proteins have been identified as SAC components, such as Mad1, Mad2, Bub1, BubR1, Bub3, Mps1, Zw10, and Rod and Aurora B kinase (reviewed by Musacchio and Salmon [35]). In vertebrates, it is believed that a diffusible inhibitory “wait anaphase” signal is generated from unattached kinetochores or lack of spindle tension (27, 45, 47) and that its primary target is Cdc20/Fzy (Fzy is the Drosophila Cdc20 homolog that we refer to as Cdc20 here), which is an essential APC/C activator (35). Mad2, BubR1 (Mad3 in Saccharomyces cerevisiae), Bub3, and Cdc20 have been found in the mitotic checkpoint complex (MCC) or its subcomplexes Bub3-BubR1-Cdc20 and Mad2-Cdc20 (42, 50, 56). Kinetochore-dependent recruitment and activation of Mad2 have been illustrated in a “template” model (12) and later a modified “two-state” model (28, 32, 35, 36, 40, 57). This model suggests that a kinetochore-bound and conformationally rearranged Mad2 is required for Cdc20 binding and that it leads to the formation of the Mad2-Cdc20 complex (8, 9, 12, 16, 48, 49). This is further supported by a more recent report that unattached kinetochores from purified HeLa cell chromosomes can catalytically generate a diffusible Cdc20 inhibitor when presented with kinetochore-bound Mad2 and that these purified chromosomes can also promote BubR1 binding to APC/C-Cdc20 by acting directly on Mad2 but not BubR1 (27). In vitro assays also suggest that Mad2 is required for Cdc20 binding to BubR1 (7, 10, 19). Fluorescence recovery after photobleaching analysis has suggested that the ∼50% of green fluorescent protein (GFP)-Cdc20 that associates with slow-phase kinetics on PtK2 cell kinetochores is Mad2 dependent (22). However, contradictory reports also exist to suggest that Mad2 might not be required for Cdc20 kinetochore localization in Xenopus and PtK2 cells (22) and that BubR1 might play a crucial role for this in human cell lines (33). In contrast to the above-mentioned slow-phase GFP-Cdc20, the remaining ∼50% of GFP-Cdc20 that associates with fast kinetics on prometaphase or metaphase kinetochores is Mad2 independent, and its kinetics parallel those of GFP-BubR1 in PtK2 cells. GFP-Cdc20 is still detectable on kinetochores through anaphase, where both Mad2 and BubR1 are greatly reduced (22, 25). Moreover, the direct requirement for the kinetochore in the formation of the SAC-inhibitory complexes has been challenged by a non-kinetochore-based formation hypothesis, with MCC found to be present in HeLa cells during S phase (50) and complex formation in yeast previously shown to be independent of intact kinetochores (17, 43). Therefore, despite the importance of Cdc20 in understanding SAC mechanisms, exactly how the SAC regulates Cdc20 via unattached kinetochores remains unclear in vertebrates.Drosophila melanogaster is a well-established model used to study the spindle assembly checkpoint (2, 3, 6, 39). More recently, phenotypes of two mad2-null Drosophila mutant alleles, mad2Δ and mad2P, have been characterized, showing that Mad2 protein is not essential for normal mitotic progression but remains essential for SAC when microtubule attachment, chromosome alignment, and congression are abnormal (5). This contrasts with its counterpart in mouse and human (14, 34, 54) and is also different from the lethality phenotypes reported for bubR1 and cdc20 mutations in Drosophila (3, 11). It has also been reported that Mad2 is less important for SAC than BubR1 and that it is regulated differently in Drosophila S2 culture cells (39). These observations led to the tentative conclusion that Drosophila Mad2 may possess different kinetochore molecular mechanisms and function differently from its homologs in mouse and human (14, 34, 54, 58). We therefore tested Mad2 kinetochore function and further investigated the mechanisms required for Cdc20 kinetochore recruitment and localization using Drosophila transgenic and mutant lines, as well as culture cells. We have characterized a new mad2-null mutant allele, mad2EY, and demonstrated that Drosophila possesses a highly conserved Mad2 kinetochore dimerization mechanism required for SAC function. However, Mad2 is not required for Cdc20 kinetochore recruitment and localization. Instead, there is an essential role for BubR1 in this mechanism during normal mitosis and SAC activation.  相似文献   

14.
Nek6 and Nek7 are members of the NIMA-related serine/threonine kinase family. Previous work showed that they contribute to mitotic progression downstream of another NIMA-related kinase, Nek9, although the roles of these different kinases remain to be defined. Here, we carried out a comprehensive analysis of the regulation and function of Nek6 and Nek7 in human cells. By generating specific antibodies, we show that both Nek6 and Nek7 are activated in mitosis and that interfering with their activity by either depletion or expression of reduced-activity mutants leads to mitotic arrest and apoptosis. Interestingly, while completely inactive mutants and small interfering RNA-mediated depletion delay cells at metaphase with fragile mitotic spindles, hypomorphic mutants or RNA interference treatment combined with a spindle assembly checkpoint inhibitor delays cells at cytokinesis. Importantly, depletion of either Nek6 or Nek7 leads to defective mitotic progression, indicating that although highly similar, they are not redundant. Indeed, while both kinases localize to spindle poles, only Nek6 obviously localizes to spindle microtubules in metaphase and anaphase and to the midbody during cytokinesis. Together, these data lead us to propose that Nek6 and Nek7 play independent roles not only in robust mitotic spindle formation but also potentially in cytokinesis.When cells divide, they must accurately segregate the duplicated genetic material between two daughter cells such that each receives a single complete set of chromosomes. This complex biomechanical feat is achieved through the action of a bipolar microtubule-based scaffold called the mitotic spindle (36). Microtubules are primarily nucleated by centrosomes that sit at the spindle poles (37). However, microtubule nucleation also occurs in the vicinity of the chromosomes and within the spindle itself (12, 13). These activities combine to ensure the efficient capture of sister chromatids as well as the maintenance of a robust structure capable of resisting the considerable forces required for chromosome separation.Spindle assembly is regulated in large part by reversible phosphorylation, and a number of protein kinases are activated during mitosis, localize to specific regions of the spindle, and phosphorylate spindle-associated proteins. These include the master mitotic regulator Cdk1/cyclin B, the polo-like kinase Plk1, and the Aurora family kinases Aurora A and B (25). More recently, members of the NIMA-related kinase family have also been implicated in mitotic spindle regulation (27, 29). NIMA was first identified in Aspergillus nidulans as a kinase required for mitotic entry, possibly through triggering the relocation of Cdk1/cyclin B to the nucleus (6, 38). NIMA can also phosphorylate S10 of histone H3 to promote chromatin condensation (7). The fission yeast NIMA-related kinase Fin1 contributes to multiple steps in mitotic progression, including the timing of mitotic entry, spindle formation, and mitotic exit (14, 15). However, the detailed mechanisms by which these fungal kinases contribute to mitotic regulation remain far from understood.In mammals, there are 11 NIMA-related kinases, named Nek1 to Nek11, and of these, 4 have been directly implicated in mitotic regulation, as follows: Nek2, Nek6, Nek7, and Nek9 (also known as Nercc1) (26, 27, 29). Nek2 is the most closely related mammalian kinase to NIMA and Fin1 by sequence and has been studied in the most detail. It localizes to the centrosome, where it phosphorylates and thereby regulates the association of a number of large coiled-coil proteins implicated in centrosome cohesion and microtubule anchoring (1, 10, 11, 21, 22, 30). These activities facilitate the early stages of spindle assembly at the G2/M transition. Interestingly, Aspergillus NIMA and fission yeast Fin1 also localize to the fungal equivalent of the centrosome, namely the spindle pole body (15, 20, 38). Here, they may participate in positive feedback loops that promote the activation of Cdk1/cyclin B and mitotic entry.Nek6, Nek7, and Nek9 act together in a mitotic kinase cascade, with Nek9 being upstream of Nek6 and Nek7. Nek9 was identified as an interacting partner of Nek6 and subsequently shown to phosphorylate Nek6 at S206 within its activation loop (2, 33). Both Nek9 and Nek6 have been reported to be activated in mitosis (2, 33, 39), although other studies dispute this (18, 23). NIMA-related kinases are characterized by having a conserved N-terminal catalytic domain, followed by a nonconserved C-terminal regulatory domain that varies in size and structure. Nek6 and Nek7 are significant exceptions to this, in that they are the smallest of the kinases and consist only of a catalytic domain with a very short N-terminal extension. They share significant similarity with each other, being 87% identical within their catalytic domains. Hence, although they exhibit distinct tissue expression patterns (8), it has generally been assumed that they are likely to have very similar properties and functions, with both being downstream substrates of Nek9.Functional studies of Nek9 reveal that it has major roles to play in the organization of the mitotic spindle. Expression of inactive and truncated Nek9 mutants led to the missegregation of chromosomes, while injection of anti-Nek9 antibodies into prophase cells caused aberrant mitotic spindle formation (33). Similarly, depletion of Nek9 from Xenopus egg extracts led to a reduction in the formation of bipolar spindles in vitro (32; J. Blot and A. M. Fry, unpublished results). The basis for these defects remains unclear, but a number of binding partners have been identified that suggest possible functions in microtubule nucleation and anchoring, including components of the γ-tubulin ring complex (γ-TuRC), the Ran GTPase, and BicD2 (18, 32, 33).While Nek9 is proposed to act upstream of Nek6 and Nek7, the proportion of its activities being channeled through these kinases is not known. Limited studies have been performed by looking at the consequences of expressing kinase-inactive Nek6 or Nek7 constructs or depleting the proteins by RNA interference (RNAi). Interference with Nek6 has been reported by one group to lead to metaphase arrest and apoptosis (39), although this is disputed by another study (23). Interference with Nek7 apparently leads to an increase in the mitotic index and apoptosis (19, 40). A decrease in centrosome-associated γ-tubulin and microtubule nucleation was also detected upon RNAi of Nek7, which is interesting in light of the interaction between Nek9 and γ-tubulin. Furthermore, defects in cytokinesis were found upon Nek7 depletion if cells were allowed to progress past the spindle checkpoint by codepletion of Mad2 (19). Importantly, both Nek9 and Nek7 localize to centrosomes, further supporting the model that this is a major site of action for this family of kinases in spindle formation (19, 32, 40).In this study, we set out to clarify the mitotic roles of Nek6 and Nek7 by examining the consequences of expression of mutants with different levels of kinase activity as well as depletion of the proteins by RNAi. Our results demonstrate that Nek6 and Nek7 are both activated in mitosis and that interference with either kinase leads to apoptosis following mitotic arrest. Interestingly, expression of inactive mutants or small interfering RNA (siRNA)-mediated depletion leads to a metaphase delay with fragile mitotic spindles, whereas expression of hypomorphic mutants or depletion in the presence of a spindle assembly checkpoint (SAC) inhibitor leads to an accumulation of cells in cytokinesis. Based on additional localization data, we propose that these kinases regulate microtubule organization not only at spindle poles but also within the mitotic spindle itself and possibly at the central spindle during late mitosis. This study therefore provides important novel insights into how Nek6 and Nek7 contribute to distinct molecular events in mitotic progression.  相似文献   

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

16.
Separase is a critical protease that catalyzes the cleavage of sister chromatid cohesins to allow the separation of sister chromatids in the anaphase. Its activity must be inhibited prior to the onset of the anaphase. Two inhibitory mechanisms exist in vertebrates that block the protease activity. One mechanism is through binding and inhibition by securin, and another is phosphorylation on Ser1126 (in humans [Ser1121 in mice]). These two mechanisms are largely redundant. However, phosphorylation on Ser1121 is critical for the prevention of premature sister separation in embryonic germ cells. As a result, Ser1121-to-Ala mutation leads to depletion of germ cells in development and subsequently to infertility in mice. Here, we report that the same mutation also causes embryogenesis failure between the 8- and 16-cell stages in mice. Our results indicate a critical role of separase phosphorylation in germ cell development as well as in early embryogenesis. Thus, deregulation of separase may be a significant contributor to infertility in humans.Sister chromatids are held together by a multisubunit complex called cohesin composed of Smc1 and -3 and Scc1 and -3 (24). To separate the sister chromosomes, cohesin complexes are removed in a two-step process. First, cohesins on chromosome arms are removed by Plk1- and Aurora B-mediated phosphorylation before the anaphase (4, 8, 19, 20, 31, 35). Second, the centromere-localized cohesins, which are protected by Sgo and PP2A from phosphorylation-mediated removal (13, 21, 28, 29, 32), are cleaved by a protease called separase at the onset of the anaphase (33, 34). Prior to the anaphase, separase is inhibited by securin and by phosphorylation, which is most likely catalyzed by cyclin B1/Cdk1. phosphorylation by cyclin B1/Cdk1 per se is not inhibitory to separase. Rather, the phosphorylation allows the binding of cyclin B1/Cdk1 as an inhibitor to separase (5). Two phosphorylation sites in separase, Ser1126 and Thr1326 (Ser1121 and Thr1321 in mice, respectively), that are important for the inhibition have been identified (30). Activation of separase depends on the function of the E3 ubiquitin ligase anaphase-promoting complex/cyclosome (APC/C), since both securin and cyclin B1 are substrates of APC/C (1-3, 15, 16, 23, 25, 27, 36). Given that APC/C is inhibited by the spindle assembly checkpoint, separation of sister chromatids therefore cannot occur until the checkpoint is satisfied. Thus, the spindle assembly checkpoint prevents premature sister separation and ensures chromosomal stability.Missegregation of chromosomes has dire consequences. It causes genetic imbalances that may transform cells and lead to cancer development in somatic tissues. In germ lines, missegregation in either meiosis I, mainly manifested as nondisjunction of homologue chromosomes, or meiosis II, manifested as premature sister chromatid separation, will generate aneuploid gametes, directly affecting the fecundity of an organism (26, 37). Although the molecular mechanisms underlying chromosome segregation errors in meiosis are still not clear (6), deregulation of separase, either directly or indirectly, is likely a significant contributor.We previously showed that securin and separase phosphorylation are redundant in almost all somatic tissues, as mice lacking either separase inhibitory mechanism are essentially normal (9, 22). However, phosphorylation of separase is uniquely required during germ line development (9). Mice carrying a nonphosphorylatable separase (S1121A) allele are sterile, largely due to depletion of germ cells during embryogenesis. The failure of the germ cells to reach sexually mature stages in the mutant mice prevented us from assessing the function of the inhibitory phosphorylation of separase in meiosis. Here we report our analysis of mice with an oocyte-specific S1121A mutation in separase. We found that these mice were still infertile. However, the infertility was not a result of meiotic errors caused by the mutant separase but was rather a failure of early embryogenesis of zygotes carrying the mutant allele prior to the 16-cell stage.  相似文献   

17.
18.
19.
In model organisms, resistance to inhibitors of cholinesterase 8 (Ric-8), a G protein α (Gα) subunit guanine nucleotide exchange factor (GEF), functions to orient mitotic spindles during asymmetric cell divisions; however, whether Ric-8A has any role in mammalian cell division is unknown. We show here that Ric-8A and Gαi function to orient the metaphase mitotic spindle of mammalian adherent cells. During mitosis, Ric-8A localized at the cell cortex, spindle poles, centromeres, central spindle, and midbody. Pertussis toxin proved to be a useful tool in these studies since it blocked the binding of Ric-8A to Gαi, thus preventing its GEF activity for Gαi. Linking Ric-8A signaling to mammalian cell division, treatment of cells with pertussis toxin, reduction of Ric-8A expression, or decreased Gαi expression similarly affected metaphase cells. Each treatment impaired the localization of LGN (GSPM2), NuMA (microtubule binding nuclear mitotic apparatus protein), and dynein at the metaphase cell cortex and disturbed integrin-dependent mitotic spindle orientation. Live cell imaging of HeLa cells expressing green fluorescent protein-tubulin also revealed that reduced Ric-8A expression prolonged mitosis, caused occasional mitotic arrest, and decreased mitotic spindle movements. These data indicate that Ric-8A signaling leads to assembly of a cortical signaling complex that functions to orient the mitotic spindle.The cortical capture of astral microtubules is essential to generate the forces needed for mitotic spindle positioning for both symmetric and asymmetric cell divisions (23, 29). Failure to either capture astral microtubules or the inappropriate application of pulling forces adversely affects mitotic spindle orientation, and can impede embryogenesis and alter cell fate decisions. Studies examining mitotic spindle orientation in Drosophila embryonic and larval neuroblasts have identified two critical pathways, the Gα/Pins/Mud pathway and the Pins/Dlg/Khc73 pathway (29). The heterotrimeric G-protein α subunit (Gα), Pins (Partner-of-Inscuteable), and Mud (Mushroom body defect) are members of an evolutionarily conserved noncanonical G-protein signaling pathway, which form a tripartite protein complex linked to the apical Par complex by the adapter protein Inscuteable (29, 37). Reducing the level of Gαi, Pins, or Mud prevents neuroblast mitotic spindle alignment. A second spindle orientation pathway involves Pins, the tumor suppressor Discs large (Dlg) and the microtubule plus-end-directed kinesin heavy chain 73 (Khc73). Khc73 binds Dlg and coimmunoprecipitates with Pins. Khc73 localized to astral microtubules can induce Pins-Dlg cortical polarity (27).In canonical G-protein signaling pathways, the binding of ligand to a seven-transmembrane receptor triggers a heterotrimeric G-protein α subunit (Gα) to exchange GTP for GDP, resulting in the dissociation of the Gα subunit from its associated Gβγ heterodimer (12, 20). This exposes interactive sites in the Gα and Gβγ subunits, allowing their binding to and activation of downstream effectors. Since Gα subunits possess an intrinsic GTPase activity, GTP hydrolysis leads to the reassembly of heterotrimeric G protein causing signaling to cease. In noncanonical G-protein signaling the seven-transmembrane receptor is replaced by an intracellular guanine nucleotide exchange factor, such as Ric-8 (37). In studies in Drosophila and Caenorhabditis elegans Ric-8 has been shown to positively regulate Gαi activity and is essential for asymmetric cell divisions (1, 2, 5, 8, 11, 36). Although initially characterized as a guanine nucleotide exchange factor (GEF) for isolated Gαsubunits, more recent biochemical studies have shown that Ric-8A (the mammalian equivalent of Ric-8) also acts on a complex of GDP-Gαi, the mammalian Pins homolog LGN, and NuMA (nuclear mitotic apparatus protein; the mammalian equivalent of Mud) catalytically releasing GTP-Gαi and causing liberation of NuMA from LGN (30, 31). Ric-8A can also catalyze guanine nucleotide exchange on Gαi1 bound to the GPR/GoLoco exchange inhibitor AGS3, a paralog of LGN (33). During mitosis the N-terminal portion of LGN binds NuMA and the C-terminal domain binds GDP-Gαi and the trimolecular complex localizes to the cell cortex, where the dynamic release of NuMA from LGN may regulate aster microtubule pulling during cell division (3, 9, 10, 22).In the present study we examined the role of Ric-8A in mitotic spindle orientation in adherent cells and in polarized MDCK cells. In nonpolarized adherent cells cell such as HeLa, integrin mediated cell-substrate adhesion orients the mitotic spindle parallel to the substratum, and thereby both daughter cells remain attached. This requires the actin cytoskeleton, astral microtubules, the microtubule plus end tracking protein EB1, myosin X, cdc42, LIM kinase 1, and phosphatidylinositol(3,4,5)-triphosphate (PIP3) (13, 18, 32, 34, 35). PIP3 may direct dynein/dynactin-dependent pulling forces on the spindle midcortex to orient the mitotic spindle (34). In polarized cells such as Madin-Darby canine kidney (MDCK) cells, the mitotic spindle is constrained by the topology of the cell and cortical cues provided by adherens junctions (24). In contrast to HeLa cells these cues are insensitive to phosphatidylinositol 3-kinase (PI3K) inhibition, which blocks the generation of PIP3 (34). We found that inhibiting either Ric-8A or Gαi expression impairs the orientation of the metaphase mitotic spindle in HeLa cells and pertussis toxin, which blocks Ric-8A triggered nucleotide exchange, disrupts the normal mitotic spindle alignment of both HeLa and MDCK cells. Impairment of Ric-8A expression or function inhibits the localization of Gαi1, LGN, NuMA, and dynein to the metaphase cortex opposite the spindle poles.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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