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.
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   

2.
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   

3.
4.
5.
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   

6.
Disruption of circadian rhythms causes decreased health and fitness, and evidence from multiple organisms links clock disruption to dysregulation of the cell cycle. However, the function of circadian regulation for the essential process of DNA replication remains elusive. Here, we demonstrate that in the cyanobacterium Synechococcus elongatus, a model organism with the simplest known circadian oscillator, the clock generates rhythms in DNA replication to minimize the number of open replication forks near dusk that would have to complete after sunset. Metabolic rhythms generated by the clock ensure that resources are available early at night to support any remaining replication forks. Combining mathematical modeling and experiments, we show that metabolic defects caused by clock–environment misalignment result in premature replisome disassembly and replicative abortion in the dark, leaving cells with incomplete chromosomes that persist through the night. Our study thus demonstrates that a major function of this ancient clock in cyanobacteria is to ensure successful completion of genome replication in a cycling environment.

Circadian clocks, internally generated rhythms in physiology with a ∼24 h period, are found in all domains of life. These clocks allow organisms to coordinate their physiological activities in anticipation of the daily cycle in the external environmental (13). Disruption of clocks caused either by mutation or clock–environment mismatch leads to decreased health and reproductive fitness in multiple organisms (46). In mammals, risk for age-related diseases such as cancer and cardiometabolic dysfunction is enhanced by circadian disruption (7, 8).Although much is now understood about the molecular mechanisms that generate rhythms, the origin of these health defects is still incompletely understood. A common target of circadian control shared across many species is the progression of the cell cycle (912). In animals, disrupted circadian rhythms are often linked to aberrant cell proliferation and tumorigenesis (13). Successful duplication of the genome is essential for the production of viable progeny. Replicating a bacterial genome can take up to several hours, a timescale over which external illumination from sunlight can change substantially. We therefore speculated that initiation of DNA replication could be a key point of circadian control. The cyanobacterium Synechococcus elongatus, which has the simplest known circadian system, is a powerful model system to address these issues, both because its clock is intimately coupled to cell cycle (9, 1416) and because clock–environment misalignment has profound effects on reproductive fitness (17). Here, we analyze whether replication is clock-regulated in S. elongatus and the consequences of clock–environment mismatch on DNA replication.  相似文献   

7.
8.
9.
DDX11 encodes an iron–sulfur cluster DNA helicase required for development, mutated, and overexpressed in cancers. Here, we show that loss of DDX11 causes replication stress and sensitizes cancer cells to DNA damaging agents, including poly ADP ribose polymerase (PARP) inhibitors and platinum drugs. We find that DDX11 helicase activity prevents chemotherapy drug hypersensitivity and accumulation of DNA damage. Mechanistically, DDX11 acts downstream of 53BP1 to mediate homology-directed repair and RAD51 focus formation in manners nonredundant with BRCA1 and BRCA2. As a result, DDX11 down-regulation aggravates the chemotherapeutic sensitivity of BRCA1/2-mutated cancers and resensitizes chemotherapy drug–resistant BRCA1/2-mutated cancer cells that regained homologous recombination proficiency. The results further indicate that DDX11 facilitates recombination repair by assisting double strand break resection and the loading of both RPA and RAD51 on single-stranded DNA substrates. We propose DDX11 as a potential target in cancers by creating pharmacologically exploitable DNA repair vulnerabilities.

Faithful DNA replication and DNA repair processes are essential for genome integrity. Inherited mutations in BRCA1 or BRCA2 genes predispose to breast and ovarian cancer, among other types of malignancies such as pancreatic cancers and brain tumors (1). Mechanistically, BRCA1 and BRCA2 are critical for double strand break (DSB) repair by homologous recombination (HR) and for the protection of stalled replication forks by facilitating RAD51 filament formation (2).Tumors with mutations in HR factors, the most widespread being those harboring mutations in BRCA1 and BRCA2, are sensitive to chemotherapeutic drugs that block replication and cause DSBs (3). Platinum drugs, such as cisplatin, create intra- and interstrand adducts that require HR activities for DNA repair during replication and therefore are effective in killing HR-defective cancers. Analysis of the plateau of the survival curve of different cancers revealed that patients often develop resistance, and thus, alternative strategies are needed. The advent of PARP (poly ADP ribose polymerase) inhibitors (PARPi), including olaparib, which exhibit synthetic lethal effects when applied to cells and tumors defective in HR (4, 5), holds significant promise. PARP1, 2, and 3 are required to repair numerous DNA single-strand breaks (SSBs) resulting from oxidative damage and during base excision repair. When PARP enzymes are locally trapped at SSBs, they prevent fork progression and generate DSBs (6), which need to be repaired by BRCA1/2 and other HR factors (4, 5). While the synthetic lethality of PARPi and HR deficiency is being exploited clinically, many BRCA-mutated carcinomas acquire resistance to PARPi (2). Identifying key factors that are functionally linked with BRCA1/2 and/or PARP during replication stress response may indicate useful alternative or combinatorial chemotherapeutic strategies.DDX11 is a conserved iron–sulfur (Fe–S) cluster 5′ to 3′ DNA helicase facilitating chromatin structure and DNA repair in manners that are not fully understood. Biallelic DDX11 mutations in humans cause the developmental disorder Warsaw breakage syndrome (WBS), which presents overlaps with Fanconi anemia in terms of chromosomal instability induced by intra- and interstrand crosslinking (ICL) agents and with cohesinopathies in terms of sister chromatid cohesion defects (7, 8). DDX11 has also strong ties to cancer. Specifically, DDX11 is highly up-regulated or amplified in diverse cancers, such as breast and ovarian cancers, including one-fifth of high-grade serous ovarian cancers (cBioPortal and The Cancer Genome Atlas [TCGA]). Moreover, DDX11 is required for the survival of advanced melanomas (9), lung adenocarcinomas (10), and hepatocellular carcinomas (11). In terms of molecular functions, DDX11 interacts physically with the replication fork component Timeless to assist replisome progression and to facilitate epigenetic stability at G-quadruplex (G4) structures and sister chromatid cohesion (1216). Notably, DDX11 also contributes along 9–1-1, Fanconi anemia factors, and SMC5/6 to prevent cytotoxicity of PARPi and ICLs (1720). However, if the DNA damage tolerance functions of DDX11 are relevant for tumorigenesis or cancer therapies remains currently unknown.Here, we find that targeting DDX11 sensitizes ovarian and other cancer cell lines to drug therapies involving cisplatin and the PARP inhibitor olaparib. We established DDX11 knockout (KO) in HeLa uterine and U2OS osteosarcoma cancer cell lines and uncovered via chemical drug screens and immunofluorescence of DNA damage markers that they show typical hallmarks of increased replication stress. DDX11 helicase activity and the Fe–S domain are critical to prevent cellular sensitization to olaparib and ICLs and to avert accumulation of DSB markers. Mechanistically, we uncover that DDX11 facilitates homology-directed repair of DSBs and RAD51 focus formation downstream of 53BP1. Importantly, DDX11 is required for viability in BRCA1-depleted cells that are resistant to chemotherapy by concomitant depletion of 53BP1, REV7, and other shieldin components (21, 22), indicating roles for DDX11 in the activated BRCA2-dependent HR pathway, often accounting for the resistance of BRCA1-mutated tumors (2). DDX11 DNA repair function is nonredundant with BRCA1 and BRCA2 pathways, facilitating resection and loading of both RPA and RAD51 on single-stranded DNA substrates. Altogether, our results define a DDX11-mediated DNA repair pathway that creates pharmaceutically targetable vulnerabilities in cancers.  相似文献   

10.
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   

11.
12.
13.
14.
Thymineless death in Escherichia coli thyA mutants growing in the absence of thymidine (dT) is preceded by a substantial resistance phase, during which the culture titer remains static, as if the chromosome has to accumulate damage before ultimately failing. Significant chromosomal replication and fragmentation during the resistance phase could provide appropriate sources of this damage. Alternatively, the initial chromosomal replication in thymine (T)-starved cells could reflect a considerable endogenous dT source, making the resistance phase a delay of acute starvation, rather than an integral part of thymineless death. Here we identify such a low-molecular-weight (LMW)-dT source as mostly dTDP-glucose and its derivatives, used to synthesize enterobacterial common antigen (ECA). The thyA mutant, in which dTDP-glucose production is blocked by the rfbA rffH mutations, lacks a LMW-dT pool, the initial DNA synthesis during T-starvation and the resistance phase. Remarkably, the thyA mutant that makes dTDP-glucose and initiates ECA synthesis normally yet cannot complete it due to the rffC defect, maintains a regular LMW-dT pool, but cannot recover dTTP from it, and thus suffers T-hyperstarvation, dying precipitously, completely losing chromosomal DNA and eventually lysing, even without chromosomal replication. At the same time, its ECA+ thyA parent does not lyse during T-starvation, while both the dramatic killing and chromosomal DNA loss in the ECA-deficient thyA mutants precede cell lysis. We conclude that: 1) the significant pool of dTDP-hexoses delays acute T-starvation; 2) T-starvation destabilizes even nonreplicating chromosomes, while T-hyperstarvation destroys them; and 3) beyond the chromosome, T-hyperstarvation also destabilizes the cell envelope.

Acute starvation for thymidine triphosphate (dTTP), one of the four precursors for DNA synthesis, is lethal in both bacterial and eukaryotic cells (1). Following a short resistance phase, the rapid death of thyA auxotrophs in media lacking thymine or thymidine (“T-starvation”) known as thymineless death (TLD) was first described in Escherichia coli (2, 3) and since then was extensively studied to identify the cause of lethality (1, 4, 5). Because the bulk of thymidine (dT) in any cell is used for chromosomal DNA synthesis, lack of dT was always assumed to cause some form of chromosomal damage, and hence the role of DNA repair pathways during T-starvation was the focus of intense investigation (69). These studies revealed that certain pathways, like double-strand break repair initiated by the RecBCD helicase/nuclease, Holliday junction resolution by RuvABC, and antirecombination activity of the UvrD helicase, keep cells alive during the resistance phase of T-starvation. Other events, like attempted single-strand gap repair initiated by the RecFOR complex, the function of the RecQ helicase and RecJ exonuclease, and SOS induction of the cell division inhibitor SulA, are detrimental for T-starved E. coli cells (8, 1012). However, the thyA mutants of E. coli inactivated for all of the latter “toxic DNA repair pathways” still die by two orders of magnitude during T-starvation (8), indicating some other yet-to-be-identified major lethality factors.Since actively growing cells continuously require a lot of dT to replicate chromosomal DNA, existing replication forks were inferred to be the points of TLD pathology (7, 8, 1315). Indeed, T-starvation severely inhibits chromosomal DNA replication (15) and is associated with accumulation of single-stranded DNA, suggesting generation of single-strand (ss) gaps by attempted replication in the absence of dT (7, 16). These ss-gaps induce the SOS response (7, 8, 17), which contributes to the pathology of TLD by induction of the SulA cell division inhibitor (8). Also, replication initiation spike in the T-starved cells triggers the destruction of the origin-centered chromosomal subdomain during TLD, suggesting that it is the demise of the nascent replication bubbles, rather than the existing replication forks, that eventually kills the chromosome (15, 17).Although the thyA mutants cannot synthesize dT, they grow normally if supplemented with exogenous dT/T. Upon removal of dT from the growth medium, the E. coli thyA strain has a two-generation-long resistance phase (also called the lag phase) (1), when the colony-forming unit (CFU) titer of the culture stays constant (Fig. 1 A, Top). This is followed by the rapid exponential death (RED) phase, when the CFU titer falls by approximately three orders of magnitude within several hours (Fig. 1 A, Top).Open in a separate windowFig. 1.A significant endogenous pool of LMW dT decreases during T-starvation. (A) The phenomenon of TLD in E. coli suggests accumulation of chromosomal damage during the resistance phase (green) that would later kill cells during the RED phase (red). The data are adapted from ref. 16. Henceforth, the data are means (n ≥ 3) ± SEM. Cultures were grown at 37 °C in the presence of dT, which was removed at time = 0, while incubation in the growth medium continued. Top, cell death begins after 1-h-long resistance phase, during which the culture titer is stable. Bottom, during the same first hour without dT, cells manage to synthesize the amount of DNA equal to half of what they already had before dT removal. However, during the RED phase genomic DNA is gradually lost. (B) Scheme of 50% methanol fractionation of the intracellular thymidine into HMW dT (the dT content of the chromosomal DNA) and LMW dT. (C) A 0.7% agarose gel analysis of the HMW and LMW fractions of the 50% methanol-treated cells, as well as pure LB treated the same way, for DNA and RNA content (staining with ethidium bromide). Inverse images of stained gels are shown, in which the indicated samples were either incubated in the buffer or with the indicated enzyme (DNase I for the top gel, RNase A for the bottom gel). (D) The size of the LMW-dT pool, normalized to the HMW-dT content of the chromosome, either during normal growth in dT-supplemented medium or during T-starvation. Thymidine was removed at time = 0. The strain is KKW58.An obvious explanation for the resistance phase is existence of an intracellular source of dT to support slow replication; however, chromosomal DNA amount was consistently reported to remain flat during TLD (15, 1820). Besides, the recent systematic test of potential candidates for a source of dT or its analogs supporting the resistance phase returned empty-handed (16). Thus, the mechanisms behind the initial resistance to T-starvation, followed by the sudden shift to the RED phase remain unclear, leading to a reasonable assumption that the resistance phase is an integral part of the TLD phenomenon, during which chromosomal damage accumulates until it becomes irreparable, ushering the RED phase (1, 5). Specific early events during the resistance phase of TLD that would later turn poisonous during the RED phase were proposed to be futile incorporation–excision cycles (1, 21), ss-gap accumulation causing the SOS induction (1, 8, 16), futile fork breakage-repair cycles (16, 22), and overinitiation from the origin (5, 15).Two recent observations, in combination with an old popular TLD explanation, further support the idea of the resistance phase as the TLD period during which chromosomal damage accumulates without affecting viability for the time being. First, the resistance phase coincides with accumulation of double-strand breaks in the chromosome, which then paradoxically disappear during the RED phase (7, 16). Second (and in contrast to the reports mentioned above of constant chromosomal DNA amount during T-starvation) (15, 1820), we have found that during the resistance phase the amount of the chromosomal DNA actually increases ∼1.5 times over the prestarvation level, but then the chromosomal DNA is apparently destabilized during the RED phase, since it is slowly reduced to the original level (16) (Fig. 1 A, Bottom). Therefore, both the apparent chromosomal replication and the significant chromosome fragmentation during the resistance phase could lead to accumulation of chromosomal damage (SOS induction is an indicator of this accumulation) (7).On the other hand, the early DNA synthesis and the resistance phase in T-starved cells could reflect the existence of a source of dT, available early on during T-starvation, that fuels the initial DNA accumulation and delays viability loss until this pool is exhausted. In other words, the resistance phase could simply postpone TLD, rather than being an integral part of it. Previously, we have tested the two obvious high-molecular-weight (HMW) dT sources, namely, the stable RNAs and the chromosomal DNA, but found that incapacitation of neither one reduced the resistance phase or precluded the early DNA synthesis during T-starvation (16). Thus, the question of whether the resistance phase is a part of the TLD phenomenon remains unresolved.In the current study, we investigated a seemingly remote possibility of a substantial low-molecular-weight (LMW)-dT pool supporting the resistance phase of T-starvation in E. coli. While the bulk of dTTP in E. coli immediately incorporates into the chromosomal DNA, a fraction of dTTP is recruited into the dTDP-hexose pool (23), to participate in the synthesis of the exopolysaccharide (EPS) capsule, made of core lipopolysaccharide (LPS) (24), O-antigen (OA) (25), and enterobacterial common antigen (ECA) in E. coli (26). The first step of this recruitment is to conjugate dTTP with glucose; the hexose moiety of the resulting dTDP-glucose then undergoes several modifications, before eventually incorporating into oligosaccharide precursors of the outer antigens, while the activating dTDP handle is released back into the DNA precursor pools (26). We ignored LMW dT before because, if the total dT content of the chromosomal DNA is taken for 100%, the pool of dTTP constitutes ∼0.7% of it, while dTDP-glucose (unresolved from other dTDP-hexoses?) adds only another 2.4% (27). No more LMW-dT species are known in the cell, so the total expected LMW-dT pool comes to ∼3% of the total dT content of the chromosomal DNA, not nearly enough to support the resistance phase with its ∼50% increase in the chromosomal DNA mass (Fig. 1A).To investigate the role of the LMW-dT pool in TLD, we started by developing a simple protocol to extract the LMW-dT pool from growing cells and to compare it to the (HMW) chromosomal dT content. Here we show that early on during T-starvation the pool of dTDP-sugars becomes the major source of dTTP for the chromosomal DNA replication. This unexpected rebalancing of the dTTP pool with the help of cell envelope metabolism delays TLD and prevents T-hyperstarvation, a significantly more lethal phenomenon accompanied by complete chromosome destruction and cell lysis.  相似文献   

15.
16.
The ammonium transporter (AMT)/methylammonium permease (MEP)/Rhesus glycoprotein (Rh) family of ammonia (NH3/NH4+) transporters has been identified in organisms from all domains of life. In animals, fundamental roles for AMT and Rh proteins in the specific transport of ammonia across biological membranes to mitigate ammonia toxicity and aid in osmoregulation, acid–base balance, and excretion have been well documented. Here, we observed enriched Amt (AeAmt1) mRNA levels within reproductive organs of the arboviral vector mosquito, Aedes aegypti, prompting us to explore the role of AMTs in reproduction. We show that AeAmt1 is localized to sperm flagella during all stages of spermiogenesis and spermatogenesis in male testes. AeAmt1 expression in sperm flagella persists in spermatozoa that navigate the female reproductive tract following insemination and are stored within the spermathecae, as well as throughout sperm migration along the spermathecal ducts during ovulation to fertilize the descending egg. We demonstrate that RNA interference (RNAi)-mediated AeAmt1 protein knockdown leads to significant reductions (∼40%) of spermatozoa stored in seminal vesicles of males, resulting in decreased egg viability when these males inseminate nonmated females. We suggest that AeAmt1 function in spermatozoa is to protect against ammonia toxicity based on our observations of high NH4+ levels in the densely packed spermathecae of mated females. The presence of AMT proteins, in addition to Rh proteins, across insect taxa may indicate a conserved function for AMTs in sperm viability and reproduction in general.

Ammonium transporters (AMTs), methylammonium permeases (MEPs), and Rhesus glycoproteins (Rh proteins) comprise a protein family with three clades, and homologs from each have been identified in virtually all domains of life (1). AMT proteins were first identified in plants (2) with the simultaneous discovery of MEP proteins in fungi (3), followed by Rh proteins in humans (4). Ammonia (NH3/NH4+) is vital for growth in plants and microorganisms and is retained in some animals for use as an osmolyte (5, 6), for buoyancy (7, 8), and for those lacking sufficient dietary nitrogen (9). In the majority of animals, however, ammonia is the toxic by-product of amino acid and nucleic acid metabolism and, accordingly, requires efficient mechanisms for its regulation, transport, and excretion (1013). AMT, MEP, and Rh proteins are responsible for the selective movement of ammonia (NH3) or ammonium (NH4+) across biological membranes, a process that all organisms require. Unlike their vertebrate, bacterial, and fungal counterparts which function as putative NH3 gas channels (1418), a myriad of evidence suggests that plant AMT proteins and closely related members in some animals are functionally distinct and facilitate electrogenic ammonium (NH4+) transport (17, 1922). In contrast to vertebrates which only possess Rh proteins (23), many invertebrates are unique in that they express both AMT and Rh proteins, sometimes in the same cell (2428). Among insects, the presence of both AMT and Rh proteins has been described in Drosophila melanogaster (29, 30) and mosquitoes that vector disease-causing pathogens, Anopheles gambiae (22, 31) and Aedes aegypti (32, 33). It is unclear whether, in these instances, AMT and Rh proteins can functionally substitute for one another, but in the anal papillae of A. aegypti larvae, knockdown of either Amt or Rh proteins causes decreases in ammonia transport, suggesting that they do not (3234). To date, studies on ammonia transporter (AMT and Rh) function in insects have focused on ammonia sensing and tasting in sensory structures (22, 30, 31, 35), ammonia detoxification and acid–base balance in muscle, digestive, and excretory organs (15, 36), and ammonia excretion in a variety of organs involved in ion and water homeostasis (9, 24, 3234).A. aegypti is the primary vector for the transmission of the human arboviral diseases Zika, yellow fever, chikungunya, and dengue virus, which are of global health concern due to rapid increases in the geographical distribution of this species, presently at its highest ever (37, 38). In light of the well-documented evolution of insecticide resistance in mosquitoes (3942), more recent methods to control disease transmission such as the sterile insect technique (43), transinfection and sterilization of mosquitoes with the bacterium Wolbachia (44), and targeted genome editing rendering adult males sterile (45) have proven effective. These methods take advantage of various aspects of mosquito reproductive biology; however, an understanding of male reproductive biology and the male contributions to female reproductive processes is still in its infancy (46). Here, we describe the expression of an A. aegypti ammonium transporter (AeAmt1) in the sperm during all stages of spermatogenesis, spermiogenesis, and egg fertilization, which is critical for fertility.  相似文献   

17.
Calcium (Ca2+)-dependent protein kinases (CDPKs or CPKs) are a unique family of Ca2+ sensor/kinase-effector proteins with diverse functions in plants. In Arabidopsis thaliana, CPK28 contributes to immune homeostasis by promoting degradation of the key immune signaling receptor-like cytoplasmic kinase BOTRYTIS-INDUCED KINASE 1 (BIK1) and additionally functions in vegetative-to-reproductive stage transition. How CPK28 controls these seemingly disparate pathways is unknown. Here, we identify a single phosphorylation site in the kinase domain of CPK28 (Ser318) that is differentially required for its function in immune homeostasis and stem elongation. We show that CPK28 undergoes intermolecular autophosphorylation on Ser318 and can additionally be transphosphorylated on this residue by BIK1. Analysis of several other phosphorylation sites demonstrates that Ser318 phosphorylation is uniquely required to prime CPK28 for Ca2+ activation at physiological concentrations of Ca2+, possibly through stabilization of the Ca2+-bound active state as indicated by intrinsic fluorescence experiments. Together, our data indicate that phosphorylation of Ser318 is required for the activation of CPK28 at low intracellular [Ca2+] to prevent initiation of an immune response in the absence of infection. By comparison, phosphorylation of Ser318 is not required for stem elongation, indicating pathway-specific requirements for phosphorylation-based Ca2+-sensitivity priming. We additionally provide evidence for a conserved function for Ser318 phosphorylation in related group IV CDPKs, which holds promise for biotechnological applications by generating CDPK alleles that enhance resistance to microbial pathogens without consequences to yield.

Protein kinases represent one of the largest eukaryotic protein superfamilies. While roughly 500 protein kinases have been identified in humans (1), the genomes of Arabidopsis thaliana (hereafter, Arabidopsis) (2) and Oryza sativa (3) encode more than 1,000 and 1,500 protein kinases, respectively, including several families unique to plants. Among these protein kinases are the receptor-like kinases (RLKs), receptor-like cytoplasmic kinases (RLCKs), and calcium-dependent protein kinases (CDPKs or CPKs) that have emerged as key regulators of plant immunity (46). Despite encompassing only 2% of most eukaryotic genomes, protein kinases phosphorylate more than 40% of cellular proteins (7, 8), reflecting their diverse roles in coordinating intracellular signaling events. Reversible phosphorylation of serine (Ser), threonine (Thr), and tyrosine (Tyr) residues can serve an array of functions including changes in protein conformation and activation state (9, 10), protein stability and degradation (11, 12), subcellular localization (1315), and interaction with protein substrates (1618).Calcium (Ca2+) is a ubiquitous secondary messenger that acts cooperatively with protein phosphorylation to propagate intracellular signals. Spatial and temporal changes in intracellular Ca2+ levels occur in response to environmental and developmental cues (1923). In plants, Ca2+ transients are decoded by four major groups of calcium sensor proteins, which possess one or more Ca2+-binding EF-hand motifs (24, 25): calmodulins (CaM), CaM-like proteins, calcineurin B–like proteins, CDPKs, and Ca2+/CaM-dependent protein kinases.At the intersection of phosphorylation cascades and Ca2+ signaling are CDPKs, a unique family of Ca2+ sensor/kinase-effector proteins. CDPKs have been identified in all land plants and green algae, as well as certain protozoan ciliates and apicomplexan parasites (26, 27). CDPKs have a conserved domain architecture, consisting of a canonical Ser/Thr protein kinase domain and an EF-hand containing Ca2+-binding CaM-like domain (CLD), linked together by an autoinhibitory junction (AIJ) and flanked by variable regions on both the amino (N) and carboxyl (C) termini (28, 29). As their name implies, most CDPKs require Ca2+ for their activation (30). Upon Ca2+ binding to all EF-hands in the CaM-like domain, a dramatic conformational change occurs, freeing the AIJ from the catalytic site of the kinase, rendering the enzyme active (3133). CDPKs vary in their sensitivity to Ca2+ (30), presumably allowing proteins to perceive distinct stimuli through differences in Ca2+-binding affinity. For example, Arabidopsis CPK4 displays half maximal kinase activity in the presence of ∼3 μM free Ca2+ (30) while CPK5 only requires ∼100 nM (34). Importantly, CDPKs are signaling hubs with documented roles in multiple distinct pathways (4, 24, 3538) and are therefore likely regulated beyond Ca2+ activation.Subfunctionalization is at least partially mediated by protein localization and interaction with pathway-specific binding partners, as is well documented for Arabidopsis CPK3 which functions in response to biotic and abiotic stimuli in distinct cellular compartments (39). Recent attention has been drawn to site-specific phosphorylation as a mechanism to regulate the activity of multifunctional kinases. For example, phosphorylation sites on the RLK BRASSINOSTEROID INSENSITIVE 1–ASSOCIATED KINASE 1 (BAK1) are differentially required for its function as a coreceptor with a subset of leucine-rich repeat –RLKs (40). Phosphoproteomic analyses indicate that CDPKs are differentially phosphorylated following exposure to distinct stimuli (4148); however, the biochemical mechanisms by which site-specific phosphorylation regulates multifunctional CDPKs is still poorly understood.Arabidopsis CPK28 is a plasma membrane–localized protein kinase with dual roles in plant immune homeostasis (4951) and phytohormone-mediated reproductive growth (52, 53). In vegetative plants, CPK28 serves as a negative regulator of immune signal amplitude by phosphorylating and activating two PLANT U-BOX–type E3 ubiquitin ligases, PUB25 and PUB26, which target the key immune RLCK BOTRYTIS-INDUCED KINASE 1 (BIK1) for proteasomal degradation (50). Owing to elevated levels of BIK1, CPK28 null plants (cpk28-1) have heightened immune responses and enhanced resistance to the bacterial pathogen Pseudomonas syringae pv. tomato DC3000 (Pto DC3000) (51). Upon transition to the reproductive stage, cpk28-1 plants additionally present shorter leaf petioles, enhanced anthocyanin production, and a reduction in stem elongation (52, 53). The molecular basis for developmental phenotypes in the cpk28-1 knockout mutant, beyond hormonal imbalance (52, 53), are comparatively unknown.Our recent work demonstrated that autophosphorylation status dictates Ca2+-sensitivity of CPK28 peptide kinase activity in vitro (54). While dephosphorylated CPK28 is stimulated by the addition of 100 μM CaCl2 compared to untreated protein, hyperphosphorylated CPK28 displayed similar levels of activity at basal Ca2+ concentrations (54). These results highlight the interesting possibility that phosphorylation status may control the activation of multifunctional kinases in distinct pathways by allowing CDPKs to respond to stimulus-specific Ca2+ signatures.In the present study, we identify a single autophosphorylation site, Ser318, that decouples the activity of CPK28 in immune signaling from its role in reproductive growth. We show that expression of a nonphosphorylatable Ser-to-Ala variant (CPK28S318A) is unable to complement the immune phenotypes of cpk28-1 mutants but is able to complement defects in stem growth. Further, we uncover a functional role for phosphorylation of Ser318 in priming CPK28 for activation at low free [Ca2+]. Together, we demonstrate that site-specific phosphorylation can direct the activity of a multifunctional kinase in distinct pathways and provide evidence for a conserved mechanism in orthologous group IV CDPKs.  相似文献   

18.
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   

19.
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   

20.
Neuropeptides and neurotrophic factors secreted from dense core vesicles (DCVs) control many brain functions, but the calcium sensors that trigger their secretion remain unknown. Here, we show that in mouse hippocampal neurons, DCV fusion is strongly and equally reduced in synaptotagmin-1 (Syt1)- or Syt7-deficient neurons, but combined Syt1/Syt7 deficiency did not reduce fusion further. Cross-rescue, expression of Syt1 in Syt7-deficient neurons, or vice versa, completely restored fusion. Hence, both sensors are rate limiting, operating in a single pathway. Overexpression of either sensor in wild-type neurons confirmed this and increased fusion. Syt1 traveled with DCVs and was present on fusing DCVs, but Syt7 supported fusion largely from other locations. Finally, the duration of single DCV fusion events was reduced in Syt1-deficient but not Syt7-deficient neurons. In conclusion, two functionally redundant calcium sensors drive neuromodulator secretion in an expression-dependent manner. In addition, Syt1 has a unique role in regulating fusion pore duration.

To date, over 100 genes encoding neuropeptides and neurotrophic factors, together referred to as neuromodulators, are identified, and most neurons express neuromodulators and neuromodulator receptors (1). Neuromodulators travel through neurons in dense core vesicles (DCVs) and, upon secretion, regulate neuronal excitability, synaptic plasticity, and neurite outgrowth (24). Dysregulation of DCV secretion is linked to many brain disorders (57). However, the molecular mechanisms that regulate neuromodulator secretion remain largely elusive.Neuromodulator secretion, like neurotransmitter secretion from synaptic vesicles (SVs), is tightly controlled by Ca2+. The Ca2+ sensors that regulate secretion have been described for other secretory pathways but not for DCV exocytosis in neurons. Synaptotagmin (Syt) and Doc2a/b are good candidate sensors due to their interaction with SNARE complexes, phospholipids, and Ca2+ (811). The Syt family consists of 17 paralogs (12, 13). Eight show Ca2+-dependent lipid binding: Syt1 to 3, Syt5 to 7, and Syt9 and 10 (14, 15). Syt1 mediates synchronous SV fusion (8), consistent with its low Ca2+-dependent lipid affinity (15, 16) and fast Ca2+/membrane dissociation kinetics (16, 17). Syt1 is also required for the fast fusion in chromaffin cells (18) and fast striatal dopamine release (19). Synaptotagmin-7 (Syt7), in contrast, drives asynchronous SV fusion (20), in line with its a higher Ca2+ affinity (15) and slower dissociation kinetics (16). Syt7 is also a major calcium sensor for neuroendocrine secretion (21) and secretion in pancreatic cells (2224). Other sensors include Syt4, which negatively regulates brain-derived neurothropic factor (25) and oxytocin release (26), in line with its Ca2+ independency. Syt9 regulates hormone secretion in the anterior pituitary (27) and, together with Syt1, secretion from PC12 cells (28, 29). Syt10 controls growth factor secretion (30). However, Syt9 and Syt10 expression is highly restricted in the brain (3133). Hence, the calcium sensors for neuronal DCV fusion remain largely elusive. Because DCVs are generally not located close to Ca2+ channels (34), we hypothesized that DCV fusion is triggered by high-affinity Ca2+ sensors. Because of their important roles in vesicle secretion, their Ca2+ binding ability, and their high expression levels in the brain (20, 31, 3538), we addressed the roles of Doc2a/b, Syt1, and Syt7 in neuronal DCV fusion.In this study, we used primary Doc2a/b-, Syt1-, and Syt7-null (knockout, KO) neurons expressing DCV fusion reporters (34, 3941) with single-vesicle resolution. We show that both Syt1 and Syt7, but not Doc2a/b, are required for ∼60 to 90% of DCV fusion events. Deficiency of both Syt1 and Syt7 did not produce an additive effect, suggesting they function in the same pathway. Syt1 overexpression (Syt1-OE) rescued DCV fusion in Syt7-null neurons, and vice versa, indicating that the two proteins compensate for each other in DCV secretion. Moreover, overexpression of Syt1 or Syt7 in wild-type (WT) neurons increased DCV fusion, suggesting they are both rate limiting for DCV secretion. We conclude that DCV fusion requires two calcium sensors, Syt1 and Syt7, that act in a single/serial pathway and that both sensors regulate fusion in a rate-limiting and dose-dependent manner.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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