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 15 毫秒
1.
2.
DNA错配修复(mismatch repair, MMR)是一种进化中保守的机制,它校正DNA复制过程中产生的错误,维持基因组的稳定性。MMR家族蛋白同时也参与多种DNA相关的生物学功能。本研究从嗜热四膜虫鉴定了一种新的错配修复蛋白MLH3基因,该基因预测编码 319 个氨基酸,在有性生殖期特异表达。免疫荧光定位表明,HA-Mlh3定位在有性生殖期减数分裂的小核和新发育的大核中。MLH3 敲除的突变体细胞株,在有性生殖发育期停滞在两大核和两小核阶段,新大核DNA复制受阻。γ-H2A.X 检测表明,新大核和小核有性生殖后期断裂的基因组不能正常修复,发育中的细胞裂解,不能形成有性生殖后代。结果表明,Mlh3参与四膜虫新大核发育过程基因组的断裂修复和复制,对四膜虫的有性生殖是必需的。  相似文献   

3.
Ubiquitous among eukaryotes, lipid droplets are organelles that function to coordinate intracellular lipid homeostasis. Their morphology and abundance is affected by numerous genes, many of which are involved in lipid metabolism. In this report we identify a Trypanosoma brucei protein kinase, LDK, and demonstrate its localization to the periphery of lipid droplets. Association with lipid droplets was abrogated when the hydrophobic domain of LDK was deleted, supporting a model in which the hydrophobic domain is associated with or inserted into the membrane monolayer of the organelle. RNA interference knockdown of LDK modestly affected the growth of mammalian bloodstream-stage parasites but did not affect the growth of insect (procyclic)-stage parasites. However, the abundance of lipid droplets dramatically decreased in both cases. This loss was dominant over treatment with myriocin or growth in delipidated serum, both of which induce lipid body biogenesis. Growth in delipidated serum also increased LDK autophosphorylation activity. Thus, LDK is required for the biogenesis or maintenance of lipid droplets and is one of the few protein kinases specifically and predominantly associated with an intracellular organelle.Trypanosoma brucei is a single-celled eukaryotic pathogen responsible for human African trypanosomiasis (also known as African sleeping sickness) and nagana in domestic animals. More than 50,000 cases of human disease occur yearly, with over 70 million people at risk. No vaccine exists, and chemotherapy is difficult to administer and prone to pathogen resistance. As T. brucei transits between the mammalian bloodstream and the tsetse fly vector during its life cycle, the organism encounters and adapts to profoundly different environmental conditions. The parasite undergoes dramatic changes in both energy (7, 51) and lipid biosynthesis and metabolism (39, 47, 49) as it shifts between these environments.Protein kinases function in numerous regulatory aspects of the cell, including control of the cell cycle and morphology, responses to stress, and transmission of signals from the extracellular environment or between compartments of the cell. As is the case in other eukaryotes, protein kinases, particularly those associated with membranes, are expected to play pivotal roles in the cell''s ability to sense and appropriately respond to its environment. Trypanosoma brucei possesses over 170 protein kinases (16, 44). Most of these can be assigned to the standard groups of protein kinases based on sequence similarity within the kinase domain. However, sequence similarities with kinases from more well-studied organisms are rarely strong enough to allow one-to-one orthologous relationships to be determined (44), and even those which appear orthologous by sequence have sometimes shown functional divergence (46). Hence, an understanding of the roles of specific protein kinases of trypanosomatids requires an individualized assessment. The initial genome analysis of the trypanosomatids (16) showed a lack of receptor tyrosine kinases, but nine T. brucei predicted serine/threonine kinases were annotated as possessing transmembrane domains. One of these was recently shown to be strategically located at a key interface between the host and parasite: the flagellar pocket (38). This eukaryotic translation initiation factor 2α (eIF2α) family kinase was postulated to play a sensory role in monitoring protein transport.Only a very small number of protein kinases of various organisms have been observed to localize to the membranes of intracellular organelles, most of them to the endoplasmic reticulum (ER) (14, 27, 50). Lipid droplets (also known as lipid bodies, adiposomes, or oil bodies in plants) are thought to arise from the ER, although the routes of protein localization to them are not well understood. They are increasingly recognized as legitimate organelles due to their dynamic roles in energy metabolism (40), lipid trafficking (41), and protection against toxic effects of nonesterified lipids and sterols (18). Studies also suggest that they function as potential protein storage depots (12) and in antigen presentation (10). Although recent efforts to expand the lipid droplet proteome have resulted in a vastly increased and in many cases surprising catalogue of potentially associated proteins (3, 5, 11, 12, 23, 37), relatively little is known as to how these structures form and are regulated within the cell.We examine here a novel T. brucei protein kinase with a predicted transmembrane domain. Surprisingly, this protein is localized intracellularly in association with lipid droplets. RNAi-mediated knockdown of this newly identified kinase, dubbed LDK (for lipid droplet kinase), reveals a role in the formation or maintenance of lipid droplets in both mammalian bloodstream-form (BF) and insect procyclic-form (PF) stages of the parasite life cycle.  相似文献   

4.
5.
The cellular and molecular cues involved in creating branched tubular networks that transport liquids or gases throughout an organism are not well understood. To identify factors required in branching and lumen formation of Drosophila tracheal terminal cells, a model for branched tubular networks, we performed a forward genetic-mosaic screen to isolate mutations affecting these processes. From this screen, we have identified the first Drosophila mutation in the gene Zpr1 (Zinc finger protein 1) by the inability of Zpr1-mutant terminal cells to form functional, gas-filled lumens. We show that Zpr1 defective cells initiate lumen formation, but are blocked from completing the maturation required for gas filling. Zpr1 is an evolutionarily conserved protein first identified in mammalian cells as a factor that binds the intracellular domain of the unactivated epidermal growth factor receptor (EGFR). We show that down-regulation of EGFR in terminal cells phenocopies Zpr1 mutations and that Zpr1 is epistatic to ectopic lumen formation driven by EGFR overexpression. However, while Zpr1 mutants are fully penetrant, defects observed when reducing EGFR activity are only partially penetrant. These results suggest that a distinct pathway operating in parallel to the EGFR pathway contributes to lumen formation, and this pathway is also dependent on Zpr1. We provide evidence that this alternative pathway may involve fibroblast growth factor receptor (FGFR) signaling. We suggest a model in which Zpr1 mediates both EGFR and FGFR signal transduction cascades required for lumen formation in terminal cells. To our knowledge, this is the first genetic evidence placing Zpr1 downstream of EGFR signaling, and the first time Zpr1 has been implicated in FGFR signaling. Finally, we show that down-regulation of Smn, a protein known to interact with Zpr1 in mammalian cells, shows defects similar to Zpr1 mutants.  相似文献   

6.
To recycle reduced sulfur to methionine in the methionine salvage pathway (MSP), 5-methylthioribulose-1-phosphate is converted to 2-keto-4-methylthiobutyrate, the methionine precursor, by four steps; dehydratase, enolase, phosphatase, and dioxygenase reactions (catalyzed by MtnB, MtnW, MtnX and MtnD, respectively, in Bacillus subtilis). It has been proposed that the MtnBD fusion enzyme in Tetrahymena thermophila catalyzes four sequential reactions from the dehydratase to dioxygenase steps, based on the results of molecular biological analyses of mutant yeast strains with knocked-out MSP genes, suggesting that new catalytic function can be acquired by fusion of enzymes. This result raises the question of how the MtnBD fusion enzyme can catalyze four very different reactions, especially since there are no homologous domains for enolase and phosphatase (MtnW and MtnX, respectively, in B. subtilis) in the peptide. Here, we tried to identify the domains responsible for catalyzing the four reactions using recombinant proteins of full-length MtnBD and each domain alone. UV-visible and 1H-NMR spectral analyses of reaction products revealed that the MtnB domain catalyzes dehydration and enolization and the MtnD domain catalyzes dioxygenation. Contrary to a previous report, conversion of 5-methylthioribulose-1-phosphate to 2-keto-4-methylthiobutyrate was dependent on addition of an exogenous phosphatase from B. subtilis. This was observed for both the MtnB domain and full-length MtnBD, suggesting that MtnBD does not catalyze the phosphatase reaction. Our results suggest that the MtnB domain of T. thermophila MtnBD acquired the new function to catalyze both the dehydratase and enolase reactions through evolutionary gene mutations, rather than fusion of MSP genes.  相似文献   

7.
The Saccharomyces cerevisiae DJP1 gene encodes a cytosolic protein homologous to Escherichia coli DnaJ. DnaJ homologues act in conjunction with molecular chaperones of the Hsp70 protein family in a variety of cellular processes. Cells with a DJP1 gene deletion are viable and exhibit a novel phenotype among cytosolic J-protein mutants in that they have a specific impairment of only one organelle, the peroxisome. The phenotype was also unique among peroxisome assembly mutants: peroxisomal matrix proteins were mislocalized to the cytoplasm to a varying extent, and peroxisomal structures failed to grow to full size and exhibited a broad range of buoyant densities. Import of marker proteins for the endoplasmic reticulum, nucleus, and mitochondria was normal. Furthermore, the metabolic adaptation to a change in carbon source, a complex multistep process, was unaffected in a DJP1 gene deletion mutant. We conclude that Djp1p is specifically required for peroxisomal protein import.  相似文献   

8.
Protein translocation across the endoplasmic reticulum membrane occurs at the Sec61 translocon. This has two essential subunits, the channel-forming multispanning membrane protein Sec61p/Sec61α and the tail-anchored Sss1p/Sec61γ, which has been proposed to “clamp” the channel. We have analyzed the function of Sss1p using a series of domain mutants and found that both the cytosolic and transmembrane clamp domains of Sss1p are essential for protein translocation. Our data reveal that the cytosolic domain is required for Sec61p interaction but that the transmembrane clamp domain is required to complete activation of the translocon after precursor targeting to Sec61p.  相似文献   

9.
10.
Photosystem I (PSI) is a large pigment-protein complex and one of the two photosystems that drive electron transfer in oxygenic photosynthesis. We identified a nuclear gene required specifically for the accumulation of PSI in a forward genetic analysis of chloroplast biogenesis in maize. This gene, designated psa2, belongs to the “GreenCut” gene set, a group of genes found in green algae and plants but not in non-photosynthetic organisms. Disruption of the psa2 ortholog in Arabidopsis likewise resulted in the specific loss of PSI proteins. PSA2 harbors a conserved domain found in DnaJ chaperones where it has been shown to form a zinc finger and to have protein-disulfide isomerase activity. Accordingly, PSA2 exhibited protein-disulfide reductase activity in vitro. PSA2 localized to the thylakoid lumen and was found in a ∼250-kDa complex harboring the peripheral PSI protein PsaG but lacking several core PSI subunits. PSA2 mRNA is coexpressed with mRNAs encoding various proteins involved in the biogenesis of the photosynthetic apparatus with peak expression preceding that of genes encoding structural components. PSA2 protein abundance was not decreased in the absence of PSI but was reduced in the absence of the PSI assembly factor Ycf3. These findings suggest that a complex harboring PSA2 and PsaG mediates thiol transactions in the thylakoid lumen that are important for the assembly of PSI.  相似文献   

11.
In many cereal crops, meiotic crossovers predominantly occur toward the ends of chromosomes and 30 to 50% of genes rarely recombine. This limits the exploitation of genetic variation by plant breeding. Previous reports demonstrate that chiasma frequency can be manipulated in plants by depletion of the synaptonemal complex protein ZIPPER1 (ZYP1) but conflict as to the direction of change, with fewer chiasmata reported in Arabidopsis thaliana and more crossovers reported for rice (Oryza sativa). Here, we use RNA interference (RNAi) to reduce the amount of ZYP1 in barley (Hordeum vulgare) to only 2 to 17% of normal zygotene levels. In the ZYP1RNAi lines, fewer than half of the chromosome pairs formed bivalents at metaphase and many univalents were observed, leading to chromosome nondisjunction and semisterility. The number of chiasmata per cell was reduced from 14 in control plants to three to four in the ZYP1-depleted lines, although the localization of residual chiasmata was not affected. DNA double-strand break formation appeared normal, but the recombination pathway was defective at later stages. A meiotic time course revealed a 12-h delay in prophase I progression to the first labeled tetrads. Barley ZYP1 appears to function similarly to ZIP1/ZYP1 in yeast and Arabidopsis, with an opposite effect on crossover number to ZEP1 in rice, another member of the Poaceae.  相似文献   

12.
13.
14.
15.
CRISPR/Cas9-mediated deletion of an Arabidopsis gene cluster encoding eight kinases supports their immunity-specific roles in sensing pathogenic effectors.

Dear Editor,ZED1-related kinases (ZRKs) associate with the nucleotide binding, Leu-rich repeat (NLR) protein HOPZ-ACTIVATED RESISTANCE1 (ZAR1) to mediate effector-triggered immunity (ETI) against at least three distinct families of pathogenic effector proteins. However, it is unknown whether ZRKs specifically function in ETI or whether they also have additional roles in immunity and/or development. Eight ZRKs are clustered in the Arabidopsis (Arabidopsis thaliana) genome, including the three members with known roles in ETI. Here, we show that an ∼14-kb CRISPR-mediated deletion of the Arabidopsis ZRK genomic cluster specifically affects ETI, with no apparent defects in pattern-recognition-receptor–triggered immunity (PTI) or development.Phytopathogens deliver effector proteins into plant cells that suppress PTI and promote the infection process (Jones and Dangl, 2006). In turn, plants have evolved NLRs that recognize effectors, leading to an ETI response. This recognition often occurs indirectly, whereby NLRs monitor host “sensor” proteins for effector-induced perturbations (Khan et al., 2016). In the absence of their respective NLRs, some of these sensors are effector virulence targets that modulate immunity and development, while others appear to be decoys that mimic virulence targets, with ETI-specific roles (van der Hoorn and Kamoun, 2008; Khan et al., 2018).The ZAR1 NLR recognizes at least six type-III secreted effector (T3SE) families from bacterial phytopathogens. This remarkable immunodiversity appears to be conveyed through associations with members of the receptor-like cytoplasmic kinase XII-2 (RLCK XII-2) family, which all display characteristics of atypical kinases (Lewis et al., 2013; Roux et al., 2014). The ZAR1-mediated ETI responses against the Pseudomonas syringae T3SEs HopZ1a and HopF1r (formerly HopF2a) require ZED1 and ZRK3, whereas recognition of the Xanthomonas campestris T3SE AvrAC requires ZRK1/RKS1 (Lewis et al., 2013; Wang et al., 2015; Seto et al., 2017). ZRKs currently have no ascribed functions outside of ZAR1-associated ETI responses and are therefore considered decoy sensors or adaptors (Lewis et al., 2013; Wang et al., 2015; Khan et al., 2018). However, functional redundancy may exist among members of the ZRK family, masking phenotypes of individual mutants beyond gene-for-gene–type ETI responses (Lewis et al., 2013). We therefore utilized the CRISPR/Cas9 system to knock out the Arabidopsis genomic region containing eight of the 13 members of the RLCK XII-2, including all ZRK genes known to contribute to ETI, to investigate any non-ETI roles of ZRKs. The 14-kb ZRK gene cluster in Arabidopsis Col-0 plants includes ZRK1, ZRK2, ZRK3, ZRK4, ZED1, ZRK6, ZRK7, and ZRK10. A CRISPR/Cas9-mediated deletion of 13.3 kb was accomplished by designing guide RNAs flanking the ends of the ZRK gene cluster, which would result in a double-stranded break on both sides of the ZRK cluster, leaving only the 5′ end 63 nucleotides (21 amino acids) of ZRK10 and the 3′ end 118 nucleotides (39 amino acids) of ZRK7 (Fig. 1A). We obtained a T1 individual (zrk_1.11) homozygous for the deletion, as well as a T1 individual heterozygous for the mutation (zrk_1.10; Fig. 1B), from which we obtained homozygous T2 (zrk_2.11) and T3 (zrk_3.10) plants, respectively. Sequencing results from zrk_2.11 confirmed that the expected region had been deleted (Supplemental Fig. S1). Plants homozygous for the ZRK gene cluster deletion were morphologically indistinguishable from wild-type Col-0 plants, as well as zar1-1 plants (Fig. 1C). In addition, zrk plant fresh weight did not significantly differ from Col-0 plants (Supplemental Fig. S2), indicating that the ZRK cluster does not play a major role in vegetative plant development.Open in a separate windowFigure 1.Deletion of the ZRK gene cluster results in loss of ZRK-mediated ETI and does not significantly alter vegetative growth. A, Representation of ZRK gene cluster before (top) and after (bottom) CRISPR/Cas9-mediated deletion depicting guide RNAs and primers used for genotyping (see Supplemental Methods S1). ZRK KO primers (magenta) were used to confirm the deletion of the ZRK cluster, while ZRK3 primers (green) were used to check if the ZRK cluster was still present in T1 individuals. B, PCR genotyping for deletion of ZRK gene cluster. Amplification of product by ZRK3 F + R primers indicates lack of deletion; amplification by ZRK KO F + R indicates deletion has occurred. Examples for wild type (WT), heterozygous (HT; zrk_1.10), and homozygous for the deletion (HM KO; zrk_1.11) are shown. T1 lines (zrk_1.10 and zrk_1.11) are compared to wild-type Col-0. C, Uninfected morphology of homozygous zrk KO plants (zrk_3.10 or zrk_2.11) compared to Col-0 and zar1-1 plants. Bar = 1 cm. D, Phenotypes of zrk_2.11 plants 7 d after being sprayed with PtoDC3000(hopZ1a; left) or PtoDC3000(hopF1r; right) relative to wild-type Col-0 and zar1-1 plants. Plant immunity and disease image-based quantification of disease symptoms is presented in Supplemental Figure S3A (Laflamme et al., 2016).Next, we wanted to confirm that the deletion of the ZRK gene cluster compromised ZRK-mediated ETI responses. We sprayed the zrk_2.11 line with PtoDC3000(hopZ1a) or PtoDC3000(hopF1r), as both T3SEs require a ZRK as well as the NLR ZAR1 for their recognition in Arabidopsis (Lewis et al., 2013; Seto et al., 2017). We observed that the zrk_2.11 line was susceptible to both PtoDC3000(hopZ1a) and PtoDC3000(hopF1r), and this susceptibility was to the same level as zar1-1 plants as quantified by plant immunity and disease image-based quantification (Fig. 1D, Supplemental Fig. S3, A and C; Laflamme et al., 2016). We observed a similar phenotype for the zrk_3.10 line, confirming that the ZRK cluster deletion compromised ZRK-mediated ETI responses (Supplemental Fig. S3, B and C). Furthermore, the ZAR1-mediated ETI responses against the P. syringae T3SEs HopBA1a, HopX1i, and HopO1c were also lost in zrk_2.11, demonstrating the ZRK-dependence of these ETI responses (Supplemental Fig. S4; Laflamme et al., 2020). To ensure that the ZRK gene cluster deletion specifically impacted ZRK-related ETI responses, the zrk_3.10 and zrk_2.11 lines were also sprayed with PtoDC3000(avrRpt2), an ETI elicitor that does not require a ZRK or ZAR1 for its recognition (Mackey et al., 2003). zrk_3.10 and zrk_2.11 plants remained resistant to PtoDC3000(avrRpt2), indicating that the ZRK gene cluster deletion specifically impacts ZRK-mediated ETI responses (Supplemental Fig. S3C). In addition, growth of virulent PtoDC3000 on the zrk_3.10 and zrk_2.11 lines was unchanged compared to wild-type Col-0 plants, indicating that the ZRKs within this cluster likely do not represent virulence targets (Supplemental Fig. S5).We then examined whether knocking out the ZRK gene cluster impacted PTI. We first measured induction of peroxidase (POX) enzyme activity, as POX enzymes are produced in response to PTI (Mott et al., 2018). After treatment with the PTI elicitor flg22, addition of the POX substrate 5-aminosalicylic acid produces a brown end-product in the presence of active POX enzymes, which is quantified by reading at an optical density of 550 nm (OD550; Mott et al., 2018). Twenty h after leaf discs were treated with flg22, zrk_3.10 and zrk_2.11 plants showed the same level of PTI-associated POX activity as wild-type Col-0 plants (Fig. 2A). To further examine the role of the ZRK gene cluster in PTI, we quantified the growth of PtoDC3000ΔhrcC, which is defective in T3SE secretion and is sensitive to altered host PTI responses under high humidity conditions such as those used in our growth assays (Guo et al., 2009; Xin et al., 2016). Growth of PtoDC3000ΔhrcC on zrk_3.10 and zrk_2.11 plants was not significantly different compared to wild-type Col-0 plants (Fig. 2B). In addition, we monitored reactive oxygen species (ROS) production and found that zrk_3.10 and zrk_2.11 plants did not show a significant difference in the ROS burst observed in wild-type Col-0 plants (Fig. 2, C and D). Finally, we treated seedlings with flg22, and found that growth of zrk_3.10 and zrk_2.11 seedlings was inhibited by the same amount as in wild-type Col-0 seedlings, indicative of a similar induction of PTI responses (Fig. 2, E and F; Gómez-Gómez et al., 1999). Together, these results indicate that the ZRK gene cluster does not play a significant role in Arabidopsis PTI responses.Open in a separate windowFigure 2.Deletion of the ZRK gene cluster does not alter pattern-recognition-receptor–triggered immune responses. A, Response to the PTI elicitor flg22 measured by POX activity. Activity from leaf discs was quantified 20 h after treatment with 1 μm of flg22 at a measurement of OD550 (n = 6; Mott et al., 2018). B, Bacterial growth of the T3SS-compromised PtoDC3000ΔhrcC on zrk KO plants (zrk_3.10 and zrk_2.11) relative to wild-type Columbia-0 (wild-type Col-0) and zar1-1 plants 3-d postinoculation. Plants were domed for the duration of the experiment (n = 8). C, Response of Col-0, zrk KO plants (zrk_3.10 and zrk_2.11), and fls2 to the PTI elicitor flg22 measured using luminol-based detection of ROS over a time course of 60 min, with relative light units measured every 2 min (n = 12). D, Boxplots of total relative light units over a period of 30 min from treatments in C (n = 12). E, Growth inhibition of seedlings 7 d after treatment with 1 μm of flg22. F, Seedling growth inhibition was quantified by measuring fresh weight of flg22-treated seedlings as a percentage of water-treated controls (n = 4). Error bars in A, B, C, D, and F, represent se. Lowercase letters represent significantly different statistical groups by Tukey’s honest significant difference test (P < 0.05). Experiments were replicated three times with similar results.Overall, our results support an ETI-specific role for ZRKs in Arabidopsis, acting as sensors of the ZAR1 NLR. Structural insights have revealed important residues required for ZAR1-ZRK1 complex formation, and these are conserved across the RLCK XII-2 family, which includes ZRKs outside the genomic cluster (Supplemental Fig. S6; Lewis et al., 2013; Wang et al., 2019). This suggests that the ZRKs outside this genomic cluster may also play a similar role as ZAR1 sensors. As such, the ZRK family would have evolved to mimic and/or interact with the numerous kinase virulence targets of pathogenic effectors, thereby expanding the surveillance potential of ZAR1.Supplemental DataThe following supplemental materials are available.
  • Supplemental Figure S1. Sequencing confirmation of the ZRK gene cluster deletion.
  • Supplemental Figure S2. Fresh weight of zrk knockout (KO) plants (zrk_3.10 and zrk_2.11) relative to wild-type Columbia-0 (wild-type Col-0) and zar1-1 plants.
  • Supplemental Figure S3. ZRK gene cluster deletion specifically compromises ZRK-dependent ETI responses.
  • Supplemental Figure S4. ZRK gene cluster compromises the ZAR1-dependent ETI responses against HopBA1a, HopO1c, and HopX1i.
  • Supplemental Figure S5. Bacterial growth of the virulent PtoDC3000 strain on zrk KO plants (zrk_3.10 and zrk_2.11) relative to wild-type Col-0 (wild-type Col-0) plants 0- and 3-d post-inoculation via syringe infiltration.
  • Supplemental Figure S6. Multiple sequence alignment of RLCK XII-2 family shows high conservation of putative ZAR1-interacting residues.
  • Supplemental Methods S1. Generation and characterization of ZRK cluster deletion lines.
  相似文献   

16.
17.
Cellular junctions are critical for intercellular communication and for the assembly of cells into tissues. Cell junctions often consist of tight junctions, which form a permeability barrier and prevent the diffusion of lipids and proteins between cell compartments, and adherens junctions, which control the adhesion of cells and link cortical actin filaments to attachment sites on the plasma membrane. Proper tight junction formation and cell polarity require the function of membrane-associated guanylate kinases (MAGUKs) that contain the PDZ protein-protein interaction domain. In contrast, less is known about how adherens junctions are assembled. Here we describe how the PDZ-containing protein DLG-1 is required for the proper formation and function of adherens junctions in Caenorhabditis elegans. DLG-1 is a MAGUK protein that is most similar in sequence to mammalian SAP97, which is found at both synapses of the CNS, as well as at cell junctions of epithelia. DLG-1 is localized to adherens junctions, and DLG-1 localization is mediated by an amino-terminal domain shared with SAP97 but not found in other MAGUK family members. DLG-1 recruits other proteins and signaling molecules to adherens junctions, while embryos that lack DLG-1 fail to recruit the proteins AJM-1 and CPI-1 to adherens junctions. DLG-1 is required for the proper organization of the actin cytoskeleton and for the morphological elongation of embryos. In contrast to other proteins that have been observed to affect adherens junction assembly and function, DLG-1 is not required to maintain cell polarity. Our results suggest a new function for MAGUK proteins distinct from their role in cell polarity.  相似文献   

18.
Congenital heart valve defects in humans occur in approximately 2% of live births and are a major source of compromised cardiac function. In this study we demonstrate that normal heart valve development and cardiac function are dependent upon Galnt1, the gene that encodes a member of the family of glycosyltransferases (GalNAc-Ts) responsible for the initiation of mucin-type O-glycosylation. In the adult mouse, compromised cardiac function that mimics human congenital heart disease, including aortic and pulmonary valve stenosis and regurgitation; altered ejection fraction; and cardiac dilation, was observed in Galnt1 null animals. The underlying phenotype is aberrant valve formation caused by increased cell proliferation within the outflow tract cushion of developing hearts, which is first detected at developmental stage E11.5. Developing valves from Galnt1 deficient animals displayed reduced levels of the proteases ADAMTS1 and ADAMTS5, decreased cleavage of the proteoglycan versican and increased levels of other extracellular matrix proteins. We also observed increased BMP and MAPK signaling. Taken together, the ablation of Galnt1 appears to disrupt the formation/remodeling of the extracellular matrix and alters conserved signaling pathways that regulate cell proliferation. Our study provides insight into the role of this conserved protein modification in cardiac valve development and may represent a new model for idiopathic valve disease.  相似文献   

19.
20.
In most cilia, the axoneme can be subdivided into three segments: proximal (the transition zone), middle (with outer doublet microtubules), and distal (with singlet extensions of outer doublet microtubules). How the functionally distinct segments of the axoneme are assembled and maintained is not well understood. DYF-1 is a highly conserved ciliary protein containing tetratricopeptide repeats. In Caenorhabditis elegans, DYF-1 is specifically needed for assembly of the distal segment (G. Ou, O. E. Blacque, J. J. Snow, M. R. Leroux, and J. M. Scholey. Nature. 436:583-587, 2005). We show that Tetrahymena cells lacking an ortholog of DYF-1, Dyf1p, can assemble only extremely short axoneme remnants that have structural defects of diverse natures, including the absence of central pair and outer doublet microtubules and incomplete or absent B tubules on the outer microtubules. Thus, in Tetrahymena, DYF-1 is needed for either assembly or stability of the entire axoneme. Our observations support the conserved function for DYF-1 in axoneme assembly or stability but also show that the consequences of loss of DYF-1 for axoneme segments are organism specific.Cilia are microtubule-rich cellular extensions that arise from basal bodies near the surfaces of most eukaryotic cell types. Defective cilia cause a wide variety of diseases, including polycystic kidney disease, primary ciliary dyskinesia, and retinal degeneration (3). A typical motile cilium has a microtubule-based framework, the axoneme, which contains nine outer (mostly doublet) microtubules and two central (singlet) microtubules. In most cilia, the axoneme can be subdivided into three segments: proximal (transition zone), middle (containing outer doublet microtubules), and distal (containing singlet extensions of peripheral microtubules). The outer doublet microtubules of the middle segment have a complete tubule A made of 13 protofilaments and an incomplete tubule B made of 11 protofilaments that is fused to the wall of the A tubule (36, 57). The outer microtubules in the distal segment lack the B tubule (32, 49). The distal segment also lacks dynein arms and radial spokes, and its microtubules are terminated by caps that are associated with the plasma membranes at the tips of cilia (11, 50). The distal segments are characterized by a high level of microtubule turnover, which could play a role in the regulation of the length of cilia (31).The mechanisms that establish the segmental subdivision of the axoneme are not well understood. Studies of Caenorhabditis elegans indicate that the distal segment is assembled using a mechanism that differs from the one utilized in the middle and proximal segments (54). In most cell types, ciliogenesis is dependent on the intraflagellar transport (IFT) pathway, a bidirectional motility of protein aggregates, known as IFT particles, that occurs along outer microtubules (10, 28, 29, 42). IFT particles are believed to provide platforms for transport of axonemal precursors (23, 44). The anterograde component of IFT that delivers cargo from the cell body to the tips of cilia is carried out by kinesin-2 motors (28, 63), whereas the cytoplasmic dynein DHC1b is responsible for the retrograde IFT (41, 43, 53). Importantly, in the well-studied amphid cilia of C. elegans, two distinct kinesin-2 complexes are involved in the anterograde IFT and differ in movement velocity: the “slow” heterotrimeric kinesin-II and the “fast” homodimeric OSM-3 kinesin (54). While kinesin-II and OSM-3 work redundantly to assemble the middle segment, OSM-3 alone functions in the assembly of the distal segment (39, 56).In C. elegans, DYF-1 is specifically required for assembly of the distal segment (39). In the DYF-1 mutant, the rate of IFT in the remaining middle segment is reduced to the level of the slow kinesin-II, suggesting that the Osm3 complex is nonfunctional and that kinesin-II functions alone in the middle segment. Thus, DYF-1 could either activate OSM-3 kinesin or dock OSM-3 to IFT particles (14, 39).However, a recent study of zebrafish has led to a different model for DYF-1 function. Zebrafish embryos that are homozygous for a loss of function of fleer, an ortholog of DYF-1, have shortened olfactory and pronephric cilia and ultrastructural defects in the axonemes. In the middle segment, the fleer axonemes have B tubules that are disconnected from the A tubule, indicating that DYF-1 functions in the middle segment and could play a role in the stability of doublet microtubules (40). Earlier, a similar mutant phenotype was reported in Tetrahymena for a mutation in the C-terminal tail domain of β-tubulin, at the glutamic acid residues that are used by posttranslational polymodifications (glycylation and glutamylation) (47). Glycylation (46) and glutamylation (12) are conserved polymeric posttranslational modifications that affect tubulin and are highly enriched on microtubules of axonemes and centrioles (reviewed in reference 20). Other studies have indicated that tubulin glutamylation contributes to the assembly and stability of axonemes and centrioles (4, 8). The fleer mutant zebrafish cilia have reduced levels of glutamylated tubulin (40). Pathak and colleagues proposed that the primary role of DYF-1/fleer is to serve as an IFT cargo adapter for a tubulin glutamic acid ligase (25) and that the effects of lack of function of DYF-1/fleer could be caused by deficiency in tubulin glutamylation in the axoneme (40). As an alternative hypothesis, the same authors proposed that DYF-1 is a structural component that stabilizes the doublet microtubules in the axoneme (40).Here, we evaluate the significance of a DYF-1 ortholog, Dyf1p, in Tetrahymena thermophila. Unexpectedly, we found that Tetrahymena cells lacking Dyf1p either fail to assemble an axoneme or can assemble an axoneme remnant. While our observations revealed major differences in the significance of DYF-1 for segmental differentiation in diverse models, it is clear that DYF-1 is a conserved and critical component that is required for assembly of the axoneme.  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