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 46 毫秒
1.
2.
Schlafen-11 (SLFN11) inactivation in ∼50% of cancer cells confers broad chemoresistance. To identify therapeutic targets and underlying molecular mechanisms for overcoming chemoresistance, we performed an unbiased genome-wide RNAi screen in SLFN11-WT and -knockout (KO) cells. We found that inactivation of Ataxia Telangiectasia- and Rad3-related (ATR), CHK1, BRCA2, and RPA1 overcome chemoresistance to camptothecin (CPT) in SLFN11-KO cells. Accordingly, we validate that clinical inhibitors of ATR (M4344 and M6620) and CHK1 (SRA737) resensitize SLFN11-KO cells to topotecan, indotecan, etoposide, cisplatin, and talazoparib. We uncover that ATR inhibition significantly increases mitotic defects along with increased CDT1 phosphorylation, which destabilizes kinetochore-microtubule attachments in SLFN11-KO cells. We also reveal a chemoresistance mechanism by which CDT1 degradation is retarded, eventually inducing replication reactivation under DNA damage in SLFN11-KO cells. In contrast, in SLFN11-expressing cells, SLFN11 promotes the degradation of CDT1 in response to CPT by binding to DDB1 of CUL4CDT2 E3 ubiquitin ligase associated with replication forks. We show that the C terminus and ATPase domain of SLFN11 are required for DDB1 binding and CDT1 degradation. Furthermore, we identify a therapy-relevant ATPase mutant (E669K) of the SLFN11 gene in human TCGA and show that the mutant contributes to chemoresistance and retarded CDT1 degradation. Taken together, our study reveals new chemotherapeutic insights on how targeting the ATR pathway overcomes chemoresistance of SLFN11-deficient cancers. It also demonstrates that SLFN11 irreversibly arrests replication by degrading CDT1 through the DDB1–CUL4CDT2 ubiquitin ligase.

Schlafen-11 (SLFN11) is an emergent restriction factor against genomic instability acting by eliminating cells with replicative damage (16) and potentially acting as a tumor suppressor (6, 7). SLFN11-expressing cancer cells are consistently hypersensitive to a broad range of chemotherapeutic drugs targeting DNA replication, including topoisomerase inhibitors, alkylating agents, DNA synthesis, and poly(ADP-ribose) polymerase (PARP) inhibitors compared to SLFN11-deficient cancer cells, which are chemoresistant (1, 2, 4, 817). Profiling SLFN11 expression is being explored for patients to predict survival and guide therapeutic choice (8, 13, 1824).The Cancer Genome Atlas (TCGA) and cancer cell databases demonstrate that SLFN11 mRNA expression is suppressed in a broad fraction of common cancer tissues and in ∼50% of all established cancer cell lines across multiple histologies (1, 2, 5, 8, 13, 25, 26). Silencing of the SLFN11 gene, like known tumor suppressor genes, is under epigenetic mechanisms through hypermethylation of its promoter region and activation of histone deacetylases (HDACs) (21, 23, 25, 26). A recent study in small-cell lung cancer patient-derived xenograft models also showed that SLFN11 gene silencing is caused by local chromatin condensation related to deposition of H3K27me3 in the gene body of SLFN11 by EZH2, a histone methyltransferase (11). Targeting epigenetic regulators is therefore an attractive combination strategy to overcome chemoresistance of SLFN11-deficient cancers (10, 25, 26). An alternative approach is to attack SLFN11-negative cancer cells by targeting the essential pathways that cells use to overcome replicative damage and replication stress. Along these lines, a prior study showed that inhibition of ATR (Ataxia Telangiectasia- and Rad3-related) kinase reverses the resistance of SLFN11-deficient cancer cells to PARP inhibitors (4). However, targeting the ATR pathway in SLFN11-deficient cells has not yet been fully explored.SLFN11 consists of two functional domains: A conserved nuclease motif in its N terminus and an ATPase motif (putative helicase) in its C terminus (2, 6). The N terminus nuclease has been implicated in the selective degradation of type II tRNAs (including those coding for ATR) and its nuclease structure can be derived from crystallographic analysis of SLFN13 whose N terminus domain is conserved with SLFN11 (27, 28). The C terminus is only present in the group III Schlafen family (24, 29). Its potential ATPase activity and relationship to chemosensitivity to DNA-damaging agents (35) imply that the ATPase/helicase of SLFN11 is involved specifically in DNA damage response (DDR) to replication stress. Indeed, inactivation of the Walker B motif of SLFN11 by the mutation E669Q suppresses SLFN11-mediated replication block (5, 30). In addition, SLFN11 contains a binding site for the single-stranded DNA binding protein RPA1 (replication protein A1) at its C terminus (3, 31) and is recruited to replication damage sites by RPA (3, 5). The putative ATPase activity of SLFN11 is not required for this recruitment (5) but is required for blocking the replication helicase complex (CMG-CDC45) and inducing chromatin accessibility at replication origins and promoter sites (5, 30). Based on these studies, our current model is that SLFN11 is recruited to “stressed” replication forks by RPA filaments formed on single-stranded DNA (ssDNA), and that the ATPase/helicase activity of SLFN11 is required for blocking replication progression and remodeling chromatin (5, 30). However, underlying mechanisms of how SLFN11 irreversibly blocks replication in DNA damage are still unclear.Increased RPA-coated ssDNA caused by DNA damage and replication fork stalling also triggers ATR kinase activation, promoting subsequent phosphorylation of CHK1, which transiently halts cell cycle progression and enables DNA repair (32). ATR inhibitors are currently in clinical development in combination with DNA replication damaging drugs (33, 34), such as topoisomerase I (TOP1) inhibitors, which are highly synergistic with ATR inhibitors in preclinical models (35). ATR inhibitors not only inhibit DNA repair, but also lead to unscheduled replication origin firing (36), which kills cancer cells (37, 38) by inducing genomic alterations due to faulty replication and mitotic catastrophe (33).The replication licensing factor CDT1 orchestrates the initiation of replication by assembling prereplication complexes (pre-RC) in G1-phase before cells enter S-phase (39). Once replication is started by loading and activation of the MCM helicase, CDT1 is degraded by the ubiquitin proteasomal pathway to prevent additional replication initiation and ensure precise genome duplication and the firing of each origin only once per cell cycle (39, 40). At the end of G2 and during mitosis, CDT1 levels rise again to control kinetochore-microtubule attachment for accurate chromosome segregation (41). Deregulated overexpression of CDT1 results in rereplication, genome instability, and tumorigenesis (42). The cellular CDT1 levels are tightly regulated by the damage-specific DNA binding protein 1 (DDB1)–CUL4CDT2 E3 ubiquitin ligase complex in G1-phase (43) and in response to DNA damage (44, 45). How CDT1 is recognized by CUL4CDT2 in response to DNA damage remains incompletely known.In the present study, starting with a human genome-wide RNAi screen, bioinformatics analyses, and mechanistic validations, we explored synthetic lethal interactions that overcome the chemoresistance of SLFN11-deficient cells to the TOP1 inhibitor camptothecin (CPT). The strongest synergistic interaction was between depletion of the ATR/CHK1-mediated DNA damage response pathways and DNA-damaging agents in SLFN11-deficient cells. We validated and expanded our molecular understanding of combinatorial strategies in SLFN11-deficient cells with the ATR (M4344 and M6620) and CHK1 (SRA737) inhibitors in clinical development (33, 46, 47) and found that ATR inhibition leads to CDT1 stabilization and hyperphosphorylation with mitotic catastrophe. Our study also establishes that SLFN11 promotes the degradation of CDT1 by binding to DDB1, an adaptor molecule of the CUL4CDT2 E3 ubiquitin ligase complex, leading to an irreversible replication block in response to replicative DNA damage.  相似文献   

3.
Genetic screens can identify synthetic lethal (SL) interactions and uncover potential anticancer therapeutic targets. However, most SL screens have utilized knockout or knockdown approaches that do not accurately mimic chemical inhibition of a target protein. Here, we test whether missense mutations can be utilized as a model for a type of protein inhibition that creates a dominant gain-of-function cytotoxicity. We expressed missense mutations in the FEN1 endonuclease and the replication-associated helicase, CHL1, that inhibited enzymatic activity but retained substrate binding, and found that these mutations elicited a dominant SL phenotype consistent with the generation of cytotoxic protein–DNA or protein–protein intermediates. Genetic screens with nuclease-defective hFEN1 and helicase-deficient yCHL1 captured dominant SL interactions, in which ectopic expression of the mutant form, in the presence of the wild-type form, caused SL in specific mutant backgrounds. Expression of nuclease-defective hFEN1 in yeast elicited DNA binding-dependent dominant SL with homologous recombination mutants. In contrast, dominant SL interactions with helicase-deficient yCHL1 were observed in spindle-associated, Ctf18-alternative replication factor C (Ctf18-RFC) clamp loader complex, and cohesin mutant backgrounds. These results highlight the different mechanisms underlying SL interactions that occur in the presence of an inhibited form of the target protein and point to the utility of modeling trapping mutations in pursuit of more clinically relevant SL interactions.

Tumor-specific genetic alterations represent vulnerabilities that can be leveraged to selectively target tumors with therapeutics (1). This can be achieved by exploiting the concept of synthetic lethality (SL), which occurs when cells carrying perturbations of two genes individually are viable but combining those perturbations results in cell lethality (2). However, while the concept of SL holds great promise, and many cancer-relevant SL genetic interactions have been identified over the past two decades, only one SL-based therapeutic has reached the clinic (3).A number of factors confound the development of SL-based therapies, such as the context dependency of genetic interactions and the complexity of overlapping functions between different complexes and pathways in the DNA damage response (3). The success rate of SL-based therapeutics could conceivably be improved by analyzing the properties of PARP inhibitors, which are the only SL-based therapeutic currently in the clinic. Inhibitors of the proteins PARP-1 and PARP-2, which are implicated in DNA repair and genome maintenance (4), were specifically developed as SL-based therapeutics for the treatment of homologous recombination (HR) repair-deficient tumors (57). Research into the mechanism of PARP SL has found that the cytotoxicity of PARP inhibitors derives not from the loss of PARP activity, per se, but rather from the trapping of PARP protein on DNA generating a PARP–DNA cytotoxic lesion. The trapped protein–DNA lesion not only generates a potentially cytotoxic lesion, it may also prevent other repair proteins from accessing the DNA damage. The PARP-DNA lesion requires BRCA-mediated HR for error-free resolution or bypass during replication. Consistent with this model, PARP inhibitors are more effective at killing BRCA-mutated cancer cells than PARP knockout or knockdown, and the cytotoxicity of PARP inhibitors correlates with their trapping ability (810).Another class of anticancer therapeutic that traps protein targets on DNA is topoisomerase inhibitors. Many topoisomerase inhibitors act as interfacial inhibitors and interfere with the catalytic cycle, preventing the resolution of a DNA–topoisomerase intermediate, thereby creating a trapped topoisomerase–DNA adduct (11). Although they were not developed as SL-based therapeutics, their efficacy is due in part to SL interactions with tumor-specific mutations affecting replication, checkpoints, or repair (12). For both PARP and topoisomerase inhibitors, the cytotoxicity derives from the formation of a toxic intermediate in the form of an inhibited protein complexed with DNA. In effect, the small molecule inhibitors when bound to their targets convert wild-type protein to a cytotoxic form that creates a “gain-of-function” DNA-damaging agent. The creation of cytotoxic protein–DrNA complexes may be generalizable to other DNA-associated proteins, and it is also possible that inhibitors may result in cytotoxic protein–protein intermediates that indirectly impact DNA processes.Most SL screens have relied on loss-of-function knockout/mutation collections, or CRISPR-knockout and RNA interference (RNAi)-based libraries that result in loss- or reduction-of-function phenotypes (3). These approaches may often not accurately model SL interactions between cancer mutations and the chemically inhibited forms of SL partner proteins. An alternative to knockout-based SL modeling is to utilize missense mutations that mimic inhibitors. Missense-derived SL genetic interactions may be more clinically relevant than interactions that are based on complete knockouts as they are assessed when the target protein is present, and retains DNA and/or protein interactions, but is inactivated. For example, specific missense mutations in Saccharomyces cerevisiae Top1 enhance the stability of the covalent topoisomerase–DNA intermediate and phenocopy the effect of the topoisomerase inhibitor camptothecin (13). These camptothecin mimetic mutations cause a dominant phenotype and have been used to screen for mutations that sensitize cells to topoisomerase trapping (14).In both DNA- and protein-trapping scenarios, the trapped inactivated protein would be predicted to elicit a dominant phenotype. This may occur when the trapped protein competes with wild-type protein for substrate and blocks the activity of the wild-type protein. In this way, dominant SL interactions can capture genetic interactions that occur in the presence of the wild-type or residual noninhibited protein. Even trapping of a small percentage of a protein target pool could catalyze a SL interaction and may provide a mechanism by which essential proteins could be exploited to elicit SL.The efficacy of PARP and topoisomerase inhibitors that trap their targets on DNA suggest that other DNA damage response-associated proteins, in particular those associated with the response to replicative stress, would be good targets for inhibition that causes trapping. The DNA flap endonuclease 1 (FEN1) and the helicase DDX11 are critical enzymes that maintain genome stability during replication and are attractive targets for the development of anticancer therapeutics. FEN1 functions in DNA replication and repair and is required for Okazaki fragment maturation. Due to its key role in DNA replication, FEN1 supports rapid proliferation of cancer cells and is overexpressed in many tumor types (15). DDX11 is a member of the iron sulfur DEAD/DEAH helicase family (16). DDX11 is up-regulated in some primary and metastatic melanomas, and suppression of DDX11 in these melanomas curtails proliferation and induces apoptosis (17). SL interaction screens, utilizing knockout mutants of the yeast ortholog rad27Δ, have identified FEN1 as a broad-spectrum target for anticancer therapeutic development (18), especially for HR-deficient tumors (19). The knockout mutant of the DDX11 yeast ortholog chl1Δ is also SL with cohesin mutations (20), which are a common genetic vulnerability in a broad range of tumors (21, 22).In proof-of-principle experiments utilizing yeast-based high-throughput genetic approaches, we used missense mutations in FEN1 and the yeast DDX11 ortholog, Chl1, to mimic a specific case of chemical inhibition in which protein activity was inhibited, but substrate binding was unaffected (23). These mutants were screened for dominant SL interactions with a panel of DNA-associated knockout mutants. Expression of these mutant proteins generated nonoverlapping dominant SL genetic interaction networks and were consistent with the formation of cytotoxic lesions that impacted DNA processes.  相似文献   

4.
Fanconi anemia (FA) is caused by defects in cellular responses to DNA crosslinking damage and replication stress. Given the constant occurrence of endogenous DNA damage and replication fork stress, it is unclear why complete deletion of FA genes does not have a major impact on cell proliferation and germ-line FA patients are able to progress through development well into their adulthood. To identify potential cellular mechanisms that compensate for the FA deficiency, we performed dropout screens in FA mutant cells with a whole genome guide RNA library. This uncovered a comprehensive genome-wide profile of FA pathway synthetic lethality, including POLI and CDK4. As little is known of the cellular function of DNA polymerase iota (Pol ι), we focused on its role in the loss-of-function FA knockout mutants. Loss of both FA pathway function and Pol ι leads to synthetic defects in cell proliferation and cell survival, and an increase in DNA damage accumulation. Furthermore, FA-deficient cells depend on the function of Pol ι to resume replication upon replication fork stalling. Our results reveal a critical role for Pol ι in DNA repair and replication fork restart and suggest Pol ι as a target for therapeutic intervention in malignancies carrying an FA gene mutation.

Fanconi anemia (FA) is a genomic instability disorder caused by biallelic or x-linked mutations in any of 22 genes. FA patients are characterized by multiple developmental abnormalities, progressive bone marrow failure, and profound cancer susceptibility (13). Germ-line FA mutations predispose an individual to breast, ovarian, pancreatic, and hematological malignancies. Somatic FA mutations have been identified in sporadic acute leukemia and breast cancer (46).The FA pathway is the major cellular mechanism responding to DNA crosslinking damage and replication stress. The 22 FA gene products fall into several functional groups. In response to DNA damage, the FANCD2/FANCI complex is monoubiquitinated, signifying the activation of the canonical FA pathway (7, 8). The monoubiquitinated FANCD2/FANCI complex most likely orchestrates the recruitment of nucleolytic factors for the processing of crosslinking DNA damage (9, 10). The FA core complex, consisting of FANCA, -B, -C, -E, -F, -G, and -M, FAAP20, FAAP24, FAAP100, and the RING domain protein FANCL, provides the E3 ligase activity for the damage-induced monoubiquitination of FANCD2/FANCI (1116). FANCP/XPF and FANCQ/SLX4, the third group of FA gene products, are nucleases or part of the nuclease scaffold, taking part in DNA cleavage for the removal of the crosslinking lesions (8, 1721). DNA double-strand breaks, as an intermediate structure of ICL (Interstrand CrossLink) repair, depend on the fourth group of FA proteins, required in homologous recombination (FANCD1/BRCA2, FANCO/RAD51C, FANCJ/BARD1, and FANCR/RAD51) (2226).In addition to the direct role in crosslinking damage repair, FA pathway components are linked to the protection of replication fork integrity during replication interruption that is not directly caused by damage to the DNA. BRCA1/2 are important in stabilizing stalled forks in an MRE11-dependent manner (27, 28). Similarly, FANCD2 and FANCI have been shown to prevent collapse of stalled replication forks (29, 30). Defects in the FA and recombination mechanisms lead to severe fork erosion and endogenous DNA damage accumulation upon reversible replication block, suggesting that the FA pathway plays a crucial role in DNA replication under both normal and perturbed growth conditions (8, 23, 3134).Given the important role of the FA pathway in replication stress, it is perplexing that cells with a completely impaired FA mechanism are capable of sustained proliferation (34, 35). Overt abnormalities are absent in mice with knockout of several key FA genes (3639). Moreover, individuals can survive without a functional FA pathway for decades (median life expectancy of 30 y for FA patients) (40). More recently, a genome-scale CRISPR-Cas9 guide RNA (gRNA) library screen has defined gene sets essential for proliferation of common model cell lines (41). None of the classic FA genes which participate in the monoubiquitination process appear to be essential in these screens. Cells deficient in classic FA genes can sustain growth despite the accumulation of endogenous DNA damage. Thus, it seems likely that compensatory mechanisms exist in FA mutant cells to support long-term viability.In this study, we sought to identify cellular mechanisms that are important for the survival of cells deficient in the FA pathway. Comparative genome-scale CRISPR/Cas9 screens were carried out in isogenic FA pathway-proficient and -deficient cells. Genes that exhibit synthetic lethality in FA mutant cells are candidates which compensate for the loss of the FA pathway function. Among the top candidates, we validated and investigated DNA polymerase (Pol) ι as a critical factor for the survival of FA mutant cells. We found that, in FA-deficient cells, Pol ι is crucial in the resumption of stressed replication forks and in suppressing the accumulation of endogenous DNA damage. This reveals a function for Pol ι in relieving DNA damage stress.  相似文献   

5.
Spontaneous deamination of DNA cytosine and adenine into uracil and hypoxanthine, respectively, causes C to T and A to G transition mutations if left unrepaired. Endonuclease Q (EndoQ) initiates the repair of these premutagenic DNA lesions in prokaryotes by cleaving the phosphodiester backbone 5′ of either uracil or hypoxanthine bases or an apurinic/apyrimidinic (AP) lesion generated by the excision of these damaged bases. To understand how EndoQ achieves selectivity toward these structurally diverse substrates without cleaving undamaged DNA, we determined the crystal structures of Pyrococcus furiosus EndoQ bound to DNA substrates containing uracil, hypoxanthine, or an AP lesion. The structures show that substrate engagement by EndoQ depends both on a highly distorted conformation of the DNA backbone, in which the target nucleotide is extruded out of the helix, and direct hydrogen bonds with the deaminated bases. A concerted swing motion of the zinc-binding and C-terminal helical domains of EndoQ toward its catalytic domain allows the enzyme to clamp down on a sharply bent DNA substrate, shaping a deep active-site pocket that accommodates the extruded deaminated base. Within this pocket, uracil and hypoxanthine bases interact with distinct sets of amino acid residues, with positioning mediated by an essential magnesium ion. The EndoQ–DNA complex structures reveal a unique mode of damaged DNA recognition and provide mechanistic insights into the initial step of DNA damage repair by the alternative excision repair pathway. Furthermore, we demonstrate that the unique activity of EndoQ is useful for studying DNA deamination and repair in mammalian systems.

Deamination of the nucleobases is one of the most common types of damage in DNA, which can result from spontaneous hydrolysis, nitrosative stress, or activities of cellular deaminase enzymes (13). The loss of the exocyclic amino group from cytosine (C) and adenine (A) bases in DNA generates uracil (U) and hypoxanthine (Hx), respectively, which mimics thymine (T) and guanine (G) in base-pairing capacity and if left unrepaired, leads to point mutations upon DNA replication (4, 5). Spontaneous deamination of cytosine is estimated to occur over 100 times per mammalian cell per day and is a significant source of mutation that accounts for approximately half of all known pathogenic single nucleotide polymorphisms in humans (69). Deamination of adenine base in DNA is slower but still occurring at 2 to 3% the rate of cytosine deamination (10, 11). Because of its sensitivity to temperature, the hydrolytic cytosine deamination is also thought to have contributed to the rapid evolution of organisms on a warm earth (12). Enzymatic cytosine deamination also contributes to human diseases; in multiple human cancers, genomic mutations introduced by the APOBEC family of single-stranded DNA cytosine deaminases play major roles in tumor evolution, promoting the development of metastases and chemotherapeutic resistance (1315).As premutagenic lesions, deaminated DNA bases are repaired by multiple cellular pathways. A highly conserved base excision repair pathway entails removal of U or Hx base from DNA by a lesion-specific N-glycosylase, followed by repair of the resulting apurinic/apyrimidinic (AP) lesion by the concerted action of DNA endonuclease, polymerase, and ligase enzymes (1618). The Uracil-DNA glycosylase (UDG/UNG) and alkyladenine DNA glycosylase excises U and Hx from DNA, respectively (19, 20). In eukaryotes, U:G mismatches generated as a result of cytosine deamination are also repaired by the mismatch repair pathway (21). In prokaryotes, endonuclease V (EndoV) cleaves the second phosphodiester bond on the 3′ side of deoxyinosine (dI: 2′-deoxynucleotide form of Hx) in double-stranded DNA, which has been proposed to initiate an alternative excision repair pathway for deaminated purine bases (2224).Recent studies by Ishino and colleagues have identified a novel enzyme from archaea and some bacteria, designated endonuclease Q (EndoQ), which exhibits a unique dual specificity for U and Hx in DNA (2527). EndoQ is structurally and mechanistically distinct from EndoV and cleaves the phosphodiester bond immediately 5′ of 2′-deoxyuridine (dU) or dI in either single-stranded or double-stranded DNA to generate 5′-phosphate and 3′-hydroxyl termini (27). In addition to DNA strands containing deaminated bases, EndoQ also cleaves the phosphodiester bond 5′ of AP lesions in DNA. However, EndoQ shows no activity toward normal (undamaged) DNA and, unlike EndoV (28), it does not recognize mismatched bases as substrates. The precise mechanisms underlying the selective recognition of deaminated pyrimidine and purine bases has remained unknown. In this study, we determined the crystal structures of EndoQ from a hyperthermophilic archaeon Pyrococcus furiosus (pfuEndoQ) in complex with dU, dI, and AP site-containing DNA substrates. These structures and supporting functional analyses combined reveal a unique mode of DNA lesion recognition by EndoQ and its utility in the studies of DNA deamination and repair in mammalian systems.  相似文献   

6.
7.
DNA end resection is a critical step in the repair of DNA double-strand breaks (DSBs) via homologous recombination (HR). However, the mechanisms governing the extent of resection at DSB sites undergoing homology-directed repair remain unclear. Here, we show that, upon DSB induction, the key resection factor CtIP is modified by the ubiquitin-like protein SUMO at lysine 578 in a PIAS4-dependent manner. CtIP SUMOylation occurs on damaged chromatin and requires prior hyperphosphorylation by the ATM protein kinase. SUMO-modified hyperphosphorylated CtIP is targeted by the SUMO-dependent E3 ubiquitin ligase RNF4 for polyubiquitination and subsequent degradation. Consequently, disruption of CtIP SUMOylation results in aberrant accumulation of CtIP at DSBs, which, in turn, causes uncontrolled excessive resection, defective HR, and increased cellular sensitivity to DSB-inducing agents. These findings reveal a previously unidentified regulatory mechanism that regulates CtIP activity at DSBs and thus the extent of end resection via ATM-dependent sequential posttranslational modification of CtIP.

DNA double-strand breaks (DSBs) constitute one of the most severe forms of DNA damage and can result in a wide variety of genetic alterations including mutations, deletions, translocations, and chromosome loss (1, 2). Extensive studies have shown that DSBs can be repaired primarily via two pathways, classical nonhomologous end joining and homologous recombination (HR), both of which are highly conserved among all eukaryotes (35). Classical nonhomologous end joining, which directly rejoins the two broken ends of a DSB, occurs throughout interphase (35). In contrast, DSB repair by HR requires the presence of a sister chromatid and is therefore restricted to the late S and G2 phases of the cell cycle (35). HR is initiated by resection of the 5′ strand of the DSB ends, yielding 3′ single-stranded DNA (ssDNA) tails that are initially coated with the replication protein A (RPA) complex (35). The resulting RPA-coated ssDNA is an essential intermediate not only in HR repair but also in ATR-CHK1 pathway activation (35). Studies conducted in yeast and mammalian cells have established that resection of DSB ends is a two-step process (35). First, the conserved MRE11/RAD50/NBS1 complex (MRE11/RAD50/XRS2 in Saccharomyces cerevisiae) cooperates with the key resection factor CtIP (Sae2 in S. cerevisiae, Ctp1 in Schizosaccharomyces pombe) to catalyze limited resection of broken DNA ends (36). Second, the resulting short 3′ overhangs are further processed through the action of either the 5′–3′ exonuclease EXO1 or the nuclease–helicase protein complex DNA2–BLM (DNA2–Sgs1 in S. cerevisiae) (35). Whereas extensive end resection is required for HR initiation and full checkpoint activation, uncontrolled and excessive processing of DSB ends can have deleterious consequences such as large deletions at DSB sites, persistent checkpoint signaling, and cell death (711). However, the mechanisms by which cells precisely control the extent of end resection at DSB sites undergoing homology-directed repair remain obscure.It has been well established that posttranslational modifications of DNA repair proteins play crucial roles in the cellular response to genotoxic stress (12, 13). For example, phosphorylation of CtIP at threonine 847 (serine 267 in Sae2) by CDK1/2 restricts its activity to the S and G2 phases of the cell cycle (1417), and promotes its capacity to stimulate the MRE11 endonuclease activity (18) as well as the annealing of broken DNA ends (19). In addition, phosphorylation of CtIP at serine 327 by CDK2 and/or Aurora A is a prerequisite for its interactions with BRCA1 and PLK1 (2023). Furthermore, CtIP undergoes phosphorylation at threonine 315 by CDK2, and this phosphorylation event regulates CtIP protein stability by facilitating its interaction with the phosphorylation-specific prolyl isomerase PIN1 (24). In addition to acting as a CDK substrate, CtIP can also be hyperphosphorylated by ATM (or ATR in Xenopus) at multiple serine/threonine–glutamine sites in response to DSBs, which is manifested by the appearance of a slow-migrating form of CtIP (25, 26). ATM-dependent hyperphosphorylation of CtIP not only facilitates its association with damaged DNA (27) but also promotes the recruitment of BLM and EXO1 to DSB sites (25). Moreover, a previous study showed that the putative ATM-targeted residues serine 231, serine 664, and serine 745 as well as the CDK-targeted residues serine 276, threonine 315, and serine 347 within CtIP are critical for its endonuclease activity, although the relative contributions of the individual modifications have not been fully characterized (28). In addition to phosphorylation, CtIP is also subject to other posttranslational modifications, such as ubiquitination and acetylation (21, 2936). Strict regulation of CtIP activity via various posttranslational modifications is crucial for accurate processing and repair of DSBs; however, precisely how these modifications are regulated in a coordinated manner remains unclear.In this study, we provide evidence that CtIP becomes SUMOylated primarily at lysine 578 upon exposure to DSB-inducing agents, and that this modification controls the activated CtIP level at DSBs and thereby the extent of DSB end resection. CtIP SUMOylation at lysine 578 is dependent on its prior hyperphosphorylation by the protein kinase ATM. SUMO-modified hyperphosphorylated CtIP can be targeted by the SUMO-dependent E3 ubiquitin ligase RNF4 for polyubiquitination and subsequent degradation. As a consequence, cells expressing non-SUMOylatable CtIP mutants exhibit aberrant accumulation of CtIP at DSB sites, uncontrolled excessive end resection, and defective HR. Our results suggest that active CtIP triggers its own SUMOylation and degradation, establishing a negative feedback loop that restricts CtIP activity at DSBs and thereby prevents excessive end resection and genome instability.  相似文献   

8.
Single-stranded DNA (ssDNA) covered with the heterotrimeric Replication Protein A (RPA) complex is a central intermediate of DNA replication and repair. How RPA is regulated to ensure the fidelity of DNA replication and repair remains poorly understood. Yeast Rtt105 is an RPA-interacting protein required for RPA nuclear import and efficient ssDNA binding. Here, we describe an important role of Rtt105 in high-fidelity DNA replication and recombination and demonstrate that these functions of Rtt105 primarily depend on its regulation of RPA. The deletion of RTT105 causes elevated spontaneous DNA mutations with large duplications or deletions mediated by microhomologies. Rtt105 is recruited to DNA double-stranded break (DSB) ends where it promotes RPA assembly and homologous recombination repair by gene conversion or break-induced replication. In contrast, Rtt105 attenuates DSB repair by the mutagenic single-strand annealing or alternative end joining pathway. Thus, Rtt105-mediated regulation of RPA promotes high-fidelity replication and recombination while suppressing repair by deleterious pathways. Finally, we show that the human RPA-interacting protein hRIP-α, a putative functional homolog of Rtt105, also stimulates RPA assembly on ssDNA, suggesting the conservation of an Rtt105-mediated mechanism.

Faithful DNA replication and repair are essential for the maintenance of genetic material (1). Even minor defects in replication or repair can cause high loads of mutations, genome instability, cancer, and other diseases (1). Deficiency in different DNA repair or replication proteins can lead to distinct mutation patterns (24). For example, deficiency in mismatch repair results in increased microsatellite instability, while deficiency in homologous recombination repair is often associated with tandem duplications or deletions (37). Sequence analysis of various cancer types has identified many distinct genome rearrangement and mutation signatures (8). However, the genetic basis for some of these signatures remains poorly understood, thus requiring further investigation in experimental models (8).In eukaryotic cells, Replication Protein A (RPA), the major single-stranded DNA (ssDNA) binding protein complex, is essential for DNA replication, repair, and recombination (913). It is also crucial for the suppression of mutations and genome instability (1417). RPA acts as a key scaffold to recruit and coordinate proteins involved in different DNA metabolic processes (14, 15, 17). As the first responder of ssDNA, RPA participates in both replication initiation and elongation (10, 12, 13). During replication or under replication stresses, the exposed ssDNA must be protected and stabilized by RPA to prevent formation of secondary structures (14, 16). RPA is also essential for DNA double-stranded break (DSB) repair by the homologous recombination (HR) pathway (1821). During HR, the 5′-terminated strands of DSBs are initially processed by the resection machinery, generating 3′-tailed ssDNA (22). The 3′-ssDNA becomes bound by the RPA complex to activate the DNA damage checkpoint (23). RPA is subsequently replaced by the Rad51 recombinase to form a Rad51 nucleoprotein filament (19, 24). This recombinase filament catalyzes invasion of the 3′-strands at the homologous sequence to form the D-loop structure, followed by repair DNA synthesis and resolution of recombination intermediates (18, 19, 24). During HR, RPA prevents the formation of DNA secondary structures and protects 3′-ssDNA from nucleolytic degradation (25). In addition, recent work implies a role of RPA in homology recognition (26).RPA is composed of three subunits, Rfa1, Rfa2, and Rfa3, and with a total of six oligonucleotide-binding (OB) motifs that mediate interactions with ssDNA or proteins (14, 17, 27). RPA can associate with ssDNA in different modes (28). It binds short DNA (8 to 10 nt) in an unstable mode and longer ssDNA (28 to 30 nt) in a high-affinity mode (2831). Recent single-molecule studies revealed that RPA binding on ssDNA is highly dynamic (28, 32). It can rapidly diffuse within the bound DNA ligand and quickly exchange between the free and ssDNA-bound states (3235). The cellular functions of RPA rely on its high ssDNA-binding affinity and its ability to interact with different proteins (28). Although RPA has a high affinity for ssDNA, recent studies have suggested that the binding of RPA on chromatin requires additional regulations (36). How RPA is regulated to ensure replication and repair fidelity remains poorly understood.Rtt105, a protein initially identified as a regulator of the Ty1 retrotransposon, has recently been shown to interact with RPA and acts as an RPA chaperone (36). It facilitates the nuclear localization of RPA and stimulates the loading of RPA at replication forks in unperturbed conditions or under replication stresses (36). Rtt105 exhibits synthetic genetic interactions with genes encoding replisome proteins and is required for heterochromatin silencing and telomere maintenance (37). The deletion of RTT105 results in increased gross chromosomal rearrangements and reduced resistance to DNA-damaging agents (36, 38). In vitro, Rtt105 can directly stimulate RPA binding to ssDNA, likely by changing the binding mode of RPA (36).In this study, by using a combination of genetic, biochemical, and single-molecule approaches, we demonstrate that Rtt105-dependent regulation of RPA promotes high-fidelity genome duplication and recombination while suppressing mutations and the low-fidelity repair pathways. We provide evidence that human hRIP-α, the putative functional homolog of yeast Rtt105, could regulate human RPA assembly on ssDNA in vitro. Our study unveils a layer of regulation on the maintenance of genome integrity that relies on dynamic RPA binding on ssDNA to ensure high-fidelity replication or recombination.  相似文献   

9.
10.
11.
Efficient and faithful replication of the genome is essential to maintain genome stability. Replication is carried out by a multiprotein complex called the replisome, which encounters numerous obstacles to its progression. Failure to bypass these obstacles results in genome instability and may facilitate errors leading to disease. Cells use accessory helicases that help the replisome bypass difficult barriers. All eukaryotes contain the accessory helicase Pif1, which tracks in a 5′–3′ direction on single-stranded DNA and plays a role in genome maintenance processes. Here, we reveal a previously unknown role for Pif1 in replication barrier bypass. We use an in vitro reconstituted Saccharomyces cerevisiae replisome to demonstrate that Pif1 enables the replisome to bypass an inactive (i.e., dead) Cas9 (dCas9) R-loop barrier. Interestingly, dCas9 R-loops targeted to either strand are bypassed with similar efficiency. Furthermore, we employed a single-molecule fluorescence visualization technique to show that Pif1 facilitates this bypass by enabling the simultaneous removal of the dCas9 protein and the R-loop. We propose that Pif1 is a general displacement helicase for replication bypass of both R-loops and protein blocks.

Efficient and faithful replication of the genome is essential to maintain genome stability and is carried out by a multiprotein complex called the replisome (14). There are numerous obstacles to progression of the replisome during the process of chromosome duplication. These obstacles include RNA-DNA hybrids (R-loops), DNA secondary structures, transcribing RNA polymerases, and other tightly bound proteins (59). Failure to bypass these barriers may result in genome instability, which can lead to cellular abnormalities and genetic disease. Cells contain various accessory helicases that help the replisome bypass these difficult barriers (1020). A subset of these helicases act on the opposite strand of the replicative helicase (1, 2, 14, 19).All eukaryotes contain an accessory helicase, Pif1, which tracks in a 5′–3′ direction on single-stranded DNA (ssDNA) (1116). Pif1 is important in pathways such as Okazaki-fragment processing and break-induced repair that require the removal of DNA-binding proteins as well as potential displacement of R-loops (1113, 21, 1518, 2225). Genetic studies and immunoprecipitation pull-down assays indicate that Pif1 interacts with PCNA (the DNA sliding clamp), Pol ε (the leading-strand polymerase), the MCMs (the motor subunits of the replicative helicase CMG), and RPA (the single-stranded DNA-binding protein) (15, 26, 27). Pif1 activity in break-induced repair strongly depends on its interaction with PCNA (26). These interactions with replisomal components suggest that Pif1 could interact with the replisome during replication. In Escherichia coli, the replicative helicase is the DnaB homohexamer that encircles the lagging strand and moves in a 5′–3′ direction (20). E. coli accessory helicases include the monomeric UvrD (helicase II) and Rep, which move in the 3′–5′ direction and operate on the opposite strand from the DnaB hexamer. It is known that these monomeric helicases promote the bypass of barriers during replication such as stalled RNA polymerases (5). The eukaryotic replicative helicase is the 11-subunit CMG (Cdc45, Mcm2–7, GINS) and tracks in the 3′–5′ direction, opposite to the direction of Pif1 (25, 28). Once activated by Mcm10, the MCM motor domains of CMG encircle the leading strand (2932). We hypothesized that, similar to UvrD and Rep in E. coli, Pif1 interacts with the replisome tracking in the opposite direction to enable bypass of replication obstacles.In this report, we use an in vitro reconstituted Saccharomyces cerevisiae replisome to study the role of Pif1 in bypass of a “dead” Cas9 (dCas9), which is a Cas9 protein that is deactivated in DNA cleavage but otherwise fully functional in DNA binding. As with Cas9, dCas9 is a single-turnover enzyme that can be programmed with a guide RNA (gRNA) to target either strand. The dCas9–gRNA complex forms a roadblock consisting of an R-loop and a tightly bound protein (dCas9), a construct that is similar to a stalled RNA polymerase. This roadblock (hereafter dCas9 R-loop) arrests replisomes independent of whether the dCas9 R-loop is targeted to the leading or lagging strand (30). Besides its utility due to its programmable nature (33), the use of the dCas9 R-loop allows us to answer several mechanistic questions. For example, the ability to program the dCas9 R-loop block to any specific sequence enables us to observe whether block removal is different depending on whether the block is on the leading or lagging strand. Furthermore, the inner diameter of CMG can accommodate double-stranded DNA (dsDNA) and possibly an R-loop, but not a dCas9 protein. Using the dCas9 R-loop block allows us to determine the fate of each of its components.Here, we report that Pif1 enables the bypass of the dCas9 R-loop by the replisome. Interestingly, dCas9 R-loops targeted to either the leading or lagging strand are bypassed with similar efficiency. In addition, the PCNA clamp is not required for bypass of the block, indicating that Pif1 does not need to interact with PCNA during bypass of the block. We used a single-molecule fluorescence imaging to show that both the dCas9 and the R-loop are displaced as an intact nucleoprotein complex. We propose that Pif1 is a general displacement helicase for replication bypass of both R-loops and protein blocks.  相似文献   

12.
We previously described a new osteogenic growth factor, osteolectin/Clec11a, which is required for the maintenance of skeletal bone mass during adulthood. Osteolectin binds to Integrin α11 (Itga11), promoting Wnt pathway activation and osteogenic differentiation by leptin receptor+ (LepR+) stromal cells in the bone marrow. Parathyroid hormone (PTH) and sclerostin inhibitor (SOSTi) are bone anabolic agents that are administered to patients with osteoporosis. Here we tested whether osteolectin mediates the effects of PTH or SOSTi on bone formation. We discovered that PTH promoted Osteolectin expression by bone marrow stromal cells within hours of administration and that PTH treatment increased serum osteolectin levels in mice and humans. Osteolectin deficiency in mice attenuated Wnt pathway activation by PTH in bone marrow stromal cells and reduced the osteogenic response to PTH in vitro and in vivo. In contrast, SOSTi did not affect serum osteolectin levels and osteolectin was not required for SOSTi-induced bone formation. Combined administration of osteolectin and PTH, but not osteolectin and SOSTi, additively increased bone volume. PTH thus promotes osteolectin expression and osteolectin mediates part of the effect of PTH on bone formation.

The maintenance and repair of the skeleton require the generation of new bone cells throughout adult life. Osteoblasts are relatively short-lived cells that are constantly regenerated, partly by skeletal stem cells within the bone marrow (1). The main source of new osteoblasts in adult bone marrow is leptin receptor-expressing (LepR+) stromal cells (24). These cells include the multipotent skeletal stem cells that give rise to the fibroblast colony-forming cells (CFU-Fs) in the bone marrow (2), as well as restricted osteogenic progenitors (5) and adipocyte progenitors (68). LepR+ cells are a major source of osteoblasts for fracture repair (2) and growth factors for hematopoietic stem cell maintenance (911).One growth factor synthesized by LepR+ cells, as well as osteoblasts and osteocytes, is osteolectin/Clec11a, a secreted glycoprotein of the C-type lectin domain superfamily (5, 12, 13). Osteolectin is an osteogenic factor that promotes the maintenance of the adult skeleton by promoting the differentiation of LepR+ cells into osteoblasts. Osteolectin acts by binding to integrin α11β1, which is selectively expressed by LepR+ cells and osteoblasts, activating the Wnt pathway (12). Deficiency for either Osteolectin or Itga11 (the gene that encodes integrin α11) reduces osteogenesis during adulthood and causes early-onset osteoporosis in mice (12, 13). Recombinant osteolectin promotes osteogenic differentiation by bone marrow stromal cells in culture and daily injection of mice with osteolectin systemically promotes bone formation.Osteoporosis is a progressive condition characterized by reduced bone mass and increased fracture risk (14). Several factors contribute to osteoporosis development, including aging, estrogen insufficiency, mechanical unloading, and prolonged glucocorticoid use (14). Existing therapies include antiresorptive agents that slow bone loss, such as bisphosphonates (15, 16) and estrogens (17), and anabolic agents that increase bone formation, such as parathyroid hormone (PTH) (18), PTH-related protein (19), and sclerostin inhibitor (SOSTi) (20). While these therapies increase bone mass and reduce fracture risk, they are not a cure.PTH promotes both anabolic and catabolic bone remodeling (2124). PTH is synthesized by the parathyroid gland and regulates serum calcium levels, partly by regulating bone formation and bone resorption (2325). PTH1R is a PTH receptor (26, 27) that is strongly expressed by LepR+ bone marrow stromal cells (8, 2830). Recombinant human PTH (Teriparatide; amino acids 1 to 34) and synthetic PTH-related protein (Abaloparatide) are approved by the US Food and Drug Administration (FDA) for the treatment of osteoporosis (19, 31). Daily (intermittent) administration of PTH increases bone mass by promoting the differentiation of osteoblast progenitors, inhibiting osteoblast and osteocyte apoptosis, and reducing sclerostin levels (3235). PTH promotes osteoblast differentiation by activating Wnt and BMP signaling in bone marrow stromal cells (28, 36, 37), although the mechanisms by which it regulates Wnt pathway activation are complex and uncertain (38).Sclerostin is a secreted glycoprotein that inhibits Wnt pathway activation by binding to LRP5/6, a widely expressed Wnt receptor (7, 8), reducing bone formation (39, 40). Sclerostin is secreted by osteocytes (8, 41), negatively regulating bone formation by inhibiting the differentiation of osteoblasts (41, 42). SOSTi (Romosozumab) is a humanized monoclonal antibody that binds sclerostin, preventing binding to LRP5/6 and increasing Wnt pathway activation and bone formation (43). It is FDA-approved for the treatment of osteoporosis (20, 44) and has activity in rodents in addition to humans (45, 46).The discovery that osteolectin is a bone-forming growth factor raises the question of whether it mediates the effects of PTH or SOSTi on osteogenesis.  相似文献   

13.
14.
15.
16.
17.
18.
19.
The extracellular matrix (ECM) provides a precise physical and molecular environment for cell maintenance, self-renewal, and differentiation in the stem cell niche. However, the nature and organization of the ECM niche is not well understood. The adult freshwater planarian Schmidtea mediterranea maintains a large population of multipotent stem cells (neoblasts), presenting an ideal model to study the role of the ECM niche in stem cell regulation. Here we tested the function of 165 planarian homologs of ECM and ECM-related genes in neoblast regulation. We identified the collagen gene family as one with differential effects in promoting or suppressing proliferation of neoblasts. col4-1, encoding a type IV collagen α-chain, had the strongest effect. RNA interference (RNAi) of col4-1 impaired tissue maintenance and regeneration, causing tissue regression. Finally, we provide evidence for an interaction between type IV collagen, the discoidin domain receptor, and neuregulin-7 (NRG-7), which constitutes a mechanism to regulate the balance of symmetric and asymmetric division of neoblasts via the NRG-7/EGFR pathway.

Across the animal kingdom, stem cell function is regulated by the microenvironment in the surrounding niche (1), where the concentration of molecular signals for self-renewal and differentiation can be precisely regulated (2). The niche affects stem cell biology in many processes, such as aging and tissue regeneration, as well as pathological conditions such as cancer (3). Most studies have been done in tissues with large stem cell populations, such as the intestinal crypt (4) and the hair follicle (5) in mice. Elucidation of the role of the stem cell niche in tissue regeneration requires the study of animals with high regenerative potential, such as freshwater planarians (flatworms) (6). Dugesia japonica and Schmidtea mediterranea are two well-studied species that possess the ability to regenerate any missing body part (6, 7).Adult S. mediterranea maintain a high number of stem cells (neoblasts)—∼10 to 30% of all somatic cells in the adult worm—with varying potency, including pluripotent cells (814). Neoblasts are the only proliferating somatic cells: they are molecularly heterogeneous, but all express piwi-1 (1518). Lineage-committed neoblasts are “progenitors” that transiently express both piwi-1 and tissue-specific genes (15, 19). Examples include early intestinal progenitors (γ neoblast, piwi-1+/hnf4+) (8, 10, 15, 1921) and early epidermal progenitors (ζ neoblast, piwi-1+/zfp-1+) (8, 15). Other progenitor markers include collagen for muscles (22), ChAT for neurons (23), and cavII for protonephridia (24, 25). During tissue regeneration, neoblasts are recruited to the wound site, where they proliferate then differentiate to replace the missing cell types (16, 26). Some neoblasts express the pluripotency marker tgs-1, and are designated as clonogenic neoblasts (cNeoblasts) (10, 11). cNeoblasts are located in the parenchymal space adjacent to the gut (11).Neoblasts are sensitive to γ-irradiation and can be preferentially depleted in the adult planarian (27). After sublethal γ-irradiation, remaining cNeoblasts can repopulate the stem cell pool within their niche (10, 11). The close proximity of neoblasts to the gut suggests gut may be a part of neoblast niche (28, 29). When gut integrity was impaired by silencing gata4/5/6, the egfr-1/nrg-1 ligand-receptor pair, or wwp1, maintenance of non–γ-neoblasts were also disrupted (20, 30, 31), but whether that indicates the gut directly regulates neoblast remains unclear. There is evidence indicating the dorsal-ventral (D/V) transverse muscles surrounding the gut may promote neoblast proliferation and migration, with the involvement of matrix metalloproteinase mt-mmpB (32, 33). The central nervous system has also been implicated in influencing neoblast maintenance through the expression of EGF homolog neuregulin-7 (nrg-7), a ligand for EGFR-3, affecting the balance of neoblast self-renewal (symmetric or asymmetric division) (34).In other model systems, an important component of the stem-cell niche is the extracellular matrix (ECM) (35). Germline stem cells in Drosophila are anchored to niche supporting cells with ECM on one side, while the opposite side is exposed to differentiation signals, allowing asymmetric cell fate outcomes for self-renewal or differentiation following division (3638). Few studies have addressed the ECM in planarians, largely due to the lack of genetic tools to manipulate the genome, the absence of antibodies to specific planarian ECM homologs, or the tools required to study cell fate changes. However, the genomes of D. japonica (3941) and S. mediterranea (4145), and single-cell RNA-sequencing (scRNA-seq) datasets for S. mediterranea are now available (11, 4650). A recent study of the planarian matrisome demonstrated that muscle cells are the primary source of many ECM proteins (51), which, together with those produced by neoblasts and supporting parenchymal cells, may constitute components of the neoblast niche. For example, megf6 and hemicentin restrict neoblast’s localization within the parenchyma (51, 52). Functional studies also implicate ECM-modifiers, such as matrix metalloproteases (MMPs) in neoblast migration and regeneration. For example, reducing the activity of the ECM-degrading enzymes mt-mmpA (26, 33), mt-mmpB (53), or mmp-1 (33) impaired neoblast migration, proliferation, or overall tissue growth, respectively. Neoblasts are also likely to interact with ECM components of the niche via cell surface receptors, such as β1 integrin, inactivation of which impairs brain regeneration (54, 55).Here, we identified planarian ECM homologs in silico, followed by systematic functional assessment of 165 ECM and ECM-related genes by RNA interference (RNAi), to determine the effect on neoblast repopulation in planarians challenged by a sublethal dose of γ-irradiation (10). Surprisingly, multiple classes of collagens were shown to have the strongest effects. In particular, we show that the type IV collagens (COLIV) of basement membranes (BMs), were required to regulate the repopulation of neoblasts as well as lineage progression to progenitor cells. Furthermore, our data support an interaction between COLIV and the discoidin domain receptor (DDR) in neurons that activates signaling of NRG-7 in the neoblasts to regulate neoblast self-renewal versus differentiation. Together, these data demonstrate multifaceted regulation of planarian stem cells by ECM components.  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