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.
It is commonly known that aldehyde dehydrogenases (ALDHs) are a promising therapeutic target in many diseases. Bui et al.—the authors of the paper I am discussing here (Biosci Rep (2021) 41(5): BSR20210491 https://doi.org/10.1042/BSR20210491)—point that there is a lack of research on the use of spices and herbs as the sources of naturally occurring modulators of ALDH activity. In order to carry out this type of research, the authors prepared ethanolic extracts of 22 spices and herbs. The main objective of the study was to investigate retinaldehyde dehydrogenases (RALDHs), of which retinal is the main substrate and ALDH2, the mitochondrial isoform, having acetaldehyde as the main substrate.The obtained results indicated that the tested extracts exhibited differential regulatory effects on RALDHs/ALDH2 and some of them showed a potential selective inhibition of the activity of RALDHs.  相似文献   

2.
The large diversity of organisms inhabiting various environmental niches on our planet are engaged in a lively exchange of biomolecules, including nutrients, hormones, and vitamins. In a quest to survive, organisms that we define as pathogens employ innovative methods to extract valuable resources from their host leading to an infection. One such instance is where plant-associated bacterial pathogens synthesize and deploy hormones or their molecular mimics to manipulate the physiology of the host plant. This commentary describes one such specific example—the mechanism of the enzyme AldA, an aldehyde dehydrogenase (ALDH) from the bacterial plant pathogen Pseudomonas syringae which produces the plant auxin hormone indole-3-acetic acid (IAA) by oxidizing the substrate indole-3-acetaldehyde (IAAld) using the cofactor nicotinamide adenine dinucleotide (NAD+) (Bioscience Reports (2020) 40(12), https://doi.org/10.1042/BSR20202959). Using mutagenesis, enzyme kinetics, and structural analysis, Zhang et al. established that the progress of the reaction hinges on the formation of two distinct conformations of NAD(H) during the reaction course. Additionally, a key mutation in the AldA active site ‘aromatic box’ changes the enzyme’s preference for an aromatic substrate to an aliphatic one. Our commentary concludes that such molecular level investigations help to establish the nature of the dynamics of NAD(H) in ALDH-catalyzed reactions, and further show that the key active site residues control substrate specificity. We also contemplate that insights from the present study can be used to engineer novel ALDH enzymes for environmental, health, and industrial applications.  相似文献   

3.
Colorectal adenocarcinoma (COAD) is one subtype of colorectal carcinoma (CRC), whose development is associated with genetics, inappropriate immune response, and environmental factors. Although significant advances have been made in the treatment of COAD, the mortality rate remains high. It is a pressing need to explore novel therapeutic targets of COAD. Available evidence indicated that immune cell infiltration was correlated with cancer prognosis. To reveal the roles of immune cells in the COAD prognosis, a study published in Bioscience Reports by Li et al. (Bioscience Reports (2021) 41, https://doi.org/10.1042/BSR20203496) analyzed data from The Cancer Genome Atlas (TCGA) and Gene Expression Omnibus (GEO) dataset. It demonstrated a beneficial effect of Th17 cells in COAD prognosis. In addition, six hub genes (KRT23, ULBP2, ASRGL1, SERPINA1, SCIN, and SLC28A2) were identified to correlate with Th17 cells and COAD prognosis, suggesting one new therapy strategy and some predictive biomarkers of COAD. These findings reported by Li et al. may pave one way to explore the molecular mechanism of COAD further.  相似文献   

4.
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   

5.
Supervised machine learning is an essential but difficult to use approach in biomedical data analysis. The Galaxy-ML toolkit (https://galaxyproject.org/community/machine-learning/) makes supervised machine learning more accessible to biomedical scientists by enabling them to perform end-to-end reproducible machine learning analyses at large scale using only a web browser. Galaxy-ML extends Galaxy (https://galaxyproject.org), a biomedical computational workbench used by tens of thousands of scientists across the world, with a suite of tools for all aspects of supervised machine learning.

This is a PLOS Computational Biology Software paper.
  相似文献   

6.
7.
8.
BackgroundWe compared protection of mycophenolate mofetil (MMF) and azathioprine (AZA) against acute cellular rejection (ACR) and chronic allograft nephropathy (CAN) in kidney transplant recipients on steroid-free, low-dose cyclosporine (CsA) microemulsion maintenance immunosuppression.Methods and findingsATHENA, a pragmatic, prospective, multicenter trial conducted by 6 Italian transplant centers, compared the outcomes of 233 consenting recipients of a first deceased donor kidney transplant induced with low-dose thymoglobulin and basiliximab and randomized to MMF (750 mg twice/day, n = 119) or AZA (75 to 125 mg/day, n = 114) added-on maintenance low-dose CsA microemulsion and 1-week steroid. In patients without acute clinical or subclinical rejections, CsA dose was progressively halved. Primary endpoint was biopsy-proven CAN. Analysis was by intention to treat.Participants were included between June 2007 and July 2012 and followed up to August 2016. Between-group donor and recipient characteristics, donor/recipient mismatches, and follow-up CsA blood levels were similar. During a median (interquartile range (IQR)) follow-up of 47.7 (44.2 to 48.9) months, 29 of 87 biopsied patients on MMF (33.3%) versus 31 of 88 on AZA (35.2%) developed CAN (hazard ratio (HR) [95% confidence interval (CI)]: 1.147 (0.691 to 1.904, p = 0.595). Twenty and 21 patients on MMF versus 34 and 14 on AZA had clinical [HR (95% CI): 0.58 (0.34 to 1.02); p = 0.057) or biopsy-proven subclinical [HR (95% CI): 1.49 (0.76 to 2.92); p = 0.249] ACR, respectively. Combined events [HR (95% CI): 0.85 (0.56 to 1.29); p = 0.438], patient and graft survival, delayed graft function (DGF), 3-year glomerular filtration rate (GFR) [53.8 (40.6;65.7) versus 49.8 (36.8;62.5) mL/min/1.73 m2, p = 0.50], and adverse events (AEs) were not significantly different between groups.Chronicity scores other than CAN predict long-term graft outcome. Study limitations include small sample size and unblinded design.ConclusionsIn this study, we found that in deceased donor kidney transplant recipients on low-dose CsA and no steroids, MMF had no significant benefits over AZA. This finding suggests that AZA, due to its lower costs, could safely replace MMF in combination with minimized immunosuppression.Trial registrationClinicalTrials.gov NCT00494741; EUDRACT 2006-005604-14.

Piero Ruggenenti and co-workers study maintenance immunosuppression in deceased-donor kidney transplantation.  相似文献   

9.
BackgroundThe dose of protamine required following cardiopulmonary bypass (CPB) is often determined by the dose of heparin required pre-CPB, expressed as a fixed ratio. Dosing based on mathematical models of heparin clearance is postulated to improve protamine dosing precision and coagulation. We hypothesised that protamine dosing based on a 2-compartment model would improve thromboelastography (TEG) parameters and reduce the dose of protamine administered, relative to a fixed ratio.Methods and findingsWe undertook a 2-stage, adaptive randomised controlled trial, allocating 228 participants to receive protamine dosed according to a mathematical model of heparin clearance or a fixed ratio of 1 mg of protamine for every 100 IU of heparin required to establish anticoagulation pre-CPB. A planned, blinded interim analysis was undertaken after the recruitment of 50% of the study cohort. Following this, the randomisation ratio was adapted from 1:1 to 1:1.33 to increase recruitment to the superior arm while maintaining study power. At the conclusion of trial recruitment, we had randomised 121 patients to the intervention arm and 107 patients to the control arm. The primary endpoint was kaolin TEG r-time measured 3 minutes after protamine administration at the end of CPB. Secondary endpoints included ratio of kaolin TEG r-time pre-CPB to the same metric following protamine administration, requirement for allogeneic red cell transfusion, intercostal catheter drainage at 4 hours postoperatively, and the requirement for reoperation due to bleeding. The trial was listed on a clinical trial registry (ClinicalTrials.gov Identifier: NCT03532594).Participants were recruited between April 2018 and August 2019. Those in the intervention/model group had a shorter mean kaolin r-time (6.58 [SD 2.50] vs. 8.08 [SD 3.98] minutes; p = 0.0016) post-CPB. The post-protamine thromboelastogram of the model group was closer to pre-CPB parameters (median pre-CPB to post-protamine kaolin r-time ratio 0.96 [IQR 0.78–1.14] vs. 0.75 [IQR 0.57–0.99]; p < 0.001). We found no evidence of a difference in median mediastinal/pleural drainage at 4 hours postoperatively (140 [IQR 75–245] vs. 135 [IQR 94–222] mL; p = 0.85) or requirement (as a binary outcome) for packed red blood cell transfusion at 24 hours postoperatively (19 [15.8%] vs. 14 [13.1%] p = 0.69). Those in the model group had a lower median protamine dose (180 [IQR 160–210] vs. 280 [IQR 250–300] mg; p < 0.001).Important limitations of this study include an unblinded design and lack of generalisability to certain populations deliberately excluded from the study (specifically children, patients with a total body weight >120 kg, and patients requiring therapeutic hypothermia to <28°C).ConclusionsUsing a mathematical model to guide protamine dosing in patients following CPB improved TEG r-time and reduced the dose administered relative to a fixed ratio. No differences were detected in postoperative mediastinal/pleural drainage or red blood cell transfusion requirement in our cohort of low-risk patients.Trial registrationClinicalTrials.gov Unique identifier NCT03532594.

Lachlan Miles and co-workers report on a randomized controlled trial seeking to optimise protamine dosing after cardiopulmonary bypass.  相似文献   

10.
Transposons are genomic parasites, and their new insertions can cause instability and spur the evolution of their host genomes. Rapid accumulation of short-read whole-genome sequencing data provides a great opportunity for studying new transposon insertions and their impacts on the host genome. Although many algorithms are available for detecting transposon insertions, the task remains challenging and existing tools are not designed for identifying de novo insertions. Here, we present a new benchmark fly dataset based on PacBio long-read sequencing and a new method TEMP2 for detecting germline insertions and measuring de novo ‘singleton’ insertion frequencies in eukaryotic genomes. TEMP2 achieves high sensitivity and precision for detecting germline insertions when compared with existing tools using both simulated data in fly and experimental data in fly and human. Furthermore, TEMP2 can accurately assess the frequencies of de novo transposon insertions even with high levels of chimeric reads in simulated datasets; such chimeric reads often occur during the construction of short-read sequencing libraries. By applying TEMP2 to published data on hybrid dysgenic flies inflicted by de-repressed P-elements, we confirmed the continuous new insertions of P-elements in dysgenic offspring before they regain piRNAs for P-element repression. TEMP2 is freely available at Github: https://github.com/weng-lab/TEMP2.  相似文献   

11.
Members of the M13 class of metalloproteases have been implicated in diseases and in reproductive fitness. Nevertheless, their physiological role remains poorly understood. To obtain a tractable model with which to analyze this protein family’s function, we characterized the gene family in Drosophila melanogaster and focused on reproductive phenotypes. The D. melanogaster genome contains 24 M13 class protease homologs, some of which are orthologs of human proteases, including neprilysin. Many are expressed in the reproductive tracts of either sex. Using RNAi we individually targeted the five Nep genes most closely related to vertebrate neprilysin, Nep1-5, to investigate their roles in reproduction. A reduction in Nep1, Nep2, or Nep4 expression in females reduced egg laying. Nep1 and Nep2 are required in the CNS and the spermathecae for wild-type fecundity. Females that are null for Nep2 also show defects as hosts of sperm competition as well as an increased rate of depletion for stored sperm. Furthermore, eggs laid by Nep2 mutant females are fertilized normally, but arrest early in embryonic development. In the male, only Nep1 was required to induce normal patterns of female egg laying. Reduction in the expression of Nep2-5 in the male did not cause any dramatic effects on reproductive fitness, which suggests that these genes are either nonessential for male fertility or perform redundant functions. Our results suggest that, consistent with the functions of neprilysins in mammals, these proteins are also required for reproduction in Drosophila, opening up this model system for further functional analysis of this protein class and their substrates.  相似文献   

12.
When DNA double-strand breaks occur, four-stranded DNA structures called Holliday junctions (HJs) form during homologous recombination. Because HJs connect homologous DNA by a covalent link, resolution of HJ is crucial to terminate homologous recombination and segregate the pair of DNA molecules faithfully. We recently identified Monokaryotic Chloroplast1 (MOC1) as a plastid DNA HJ resolvase in algae and plants. Although Cruciform cutting endonuclease1 (CCE1) was identified as a mitochondrial DNA HJ resolvase in yeasts, homologs or other mitochondrial HJ resolvases have not been identified in other eukaryotes. Here, we demonstrate that MOC1 depletion in the green alga Chlamydomonas reinhardtii and the moss Physcomitrella patens induced ectopic recombination between short dispersed repeats in ptDNA. In addition, MOC1 depletion disorganized thylakoid membranes in plastids. In some land plant lineages, such as the moss P. patens, a liverwort and a fern, MOC1 dually targeted to plastids and mitochondria. Moreover, mitochondrial targeting of MOC1 was also predicted in charophyte algae and some land plant species. Besides causing instability of plastid DNA, MOC1 depletion in P. patens induced short dispersed repeat-mediated ectopic recombination in mitochondrial DNA and disorganized cristae in mitochondria. Similar phenotypes in plastids and mitochondria were previously observed in mutants of plastid-targeted (RECA2) and mitochondrion-targeted (RECA1) recombinases, respectively. These results suggest that MOC1 functions in the double-strand break repair in which a recombinase generates HJs and MOC1 resolves HJs in mitochondria of some lineages of algae and plants as well as in plastids in algae and plants.

Mitochondria and plastids were established in eukaryotic cells by endosymbiotic events of α-proteobacterial and cyanobacterial ancestors, respectively (Gray, 1992; Archibald, 2015). Reminiscent of their bacterial ancestors, both organelles possess their own genomes and proliferate by division of preexisting ones (Martin and Kowallik, 1999). Plastid DNA (ptDNA) and mitochondrial DNA (mtDNA) encode some components of the photosynthetic apparatus and respiratory chain, respectively (Allen, 2003). Thus, to maintain the functions of plastids and mitochondria, ptDNA and mtDNA must faithfully replicate and segregate during proliferation of the organelles.The mitochondrion and plastid possess multiple copies of DNA, which are organized with proteins into nucleoids (Kuroiwa, 1991; Pfalz and Pfannschmidt, 2013). Nucleoids, which can be visualized as dot-like or globular structures in mitochondria and plastids when stained with DNA-specific fluorochromes such as 4′, 6-diamidino-2-phenylindole (DAPI) or SYBR Green I, are ubiquitously observed in diverse lineages of algae and plants (Kuroiwa, 1991; Sato, 2001; Kobayashi et al., 2016). The morphology of nucleoids dynamically changes according to cell cycle progression and development (Powikrowska et al., 2014). For example, in the unicellular green alga Chlamydomonas reinhardtii, ∼80 copies of ptDNA are packaged into 5 to 8 globular nucleoids in a single cup-shaped plastid during the gap 1 (G1) phase (Armbrust, 1998). Prior to plastid division, during the synthesis (S) and mitosis (M) phases, plastid nucleoids change into filamentous structures and are scattered throughout the plastids. Then the nucleoids are inherited by two daughter plastids stochastically (Ehara et al., 1990; Kamimura et al., 2018). A similar morphological change of plastid nucleoids is also observed in the plant Arabidopsis (Arabidopsis thaliana; Terasawa and Sato, 2005), and thus, the mechanism of nucleoid segregation is apparently conserved in algae and plants. However, the molecular mechanisms underlying organelle DNA segregation and changes in nucleoid morphology have remained largely unknown.In a previous study using the green alga C. reinhardtii, mutants defective in nucleoid segregation were isolated (Misumi et al., 1999). One of the mutants possessed only a single enlarged nucleoid in a plastid (Fig. 1A), which was inherited by daughter plastids unevenly (Misumi et al., 1999). Later, the mutation responsible for this phenotype was identified in a previously uncharacterized gene, Monokaryotic Chloroplast1 (MOC1), which is conserved in eukaryotic algae and plants (Kobayashi et al., 2017). MOC1 exhibited endonuclease activity in vitro, where it specifically cleaved Holliday junctions (HJs), four-stranded DNA structures formed during homologous recombination (HR; Kobayashi et al., 2017). MOC1 symmetrically introduced nicks between consecutive cytosines (C↓C, where the arrow indicates the cleavage point) at the core of HJs (Kobayashi et al., 2017). Because HJs provide a covalent link between recombining DNA molecules and must be removed prior to genome segregation (Liu and West, 2004; West, 2009), it was suggested that HR affects the nucleoid structure and MOC1 segregates plastid nucleoids by cleaving HJs between ptDNA molecules prior to plastid division (Kobayashi et al., 2017).Open in a separate windowFigure 1.MOC1 depletion destabilizes ptDNA by increasing ectopic recombination between SDRs in C. reinhardtii. A, Differential interference contrast (DIC) and fluorescent images of SYBR Green I-stained wild-type (WT) and MOC1 KO cells. Green is SYBR Green I-stained DNA and magenta is plastid chlorophyll fluorescence (Chl). The arrow indicates a plastic nucleoid. N, Nucleus. Scale bars = 5 μm. B, qPCR comparing relative copy number of ptDNA between the wild type and CrMOC1 KO. The values of plastid rpl2, psbB, chlN, and psbD loci were normalized with that of the nuclear CBLP locus. For CrMOC1 KO, two independent clones (clones 1 and 2) were analyzed. The normalized value of the wild type was defined as 1.0. The error bar represents the mean ± sd (n = 3). Asterisks indicate significant difference by Student’s t test (**P < 0.01). C, Positions of SDRs CD5, CD5′, CI12, and CD15 in C. reinhardtii ptDNA (Odahara et al., 2016). Large inverted repeats (IRa and IRb) are shown by bold lines. SDRs are indicated by triangles. D to F, qPCR comparing relative copy numbers of ectopic recombinants of CD5 (D), CI12 (E), and CD15 (F). The recombinants were quantified with primers designed as shown in Supplemental Figure S2. Each value was normalized with that of the plastid psbB locus. The error bar represents the mean ± sd (n = 3). Asterisks indicate significant difference by Student’s t test (ns, P ≥ 0.05, *P < 0.05, and **P < 0.01).In general, HR follows either of two main pathways, the double-strand break repair (DSBR) or the synthesis-dependent strand annealing (SDSA) pathway (Supplemental Fig. S1; Holliday, 1964; Szostak et al., 1983; Pâques and Haber, 1999). The two pathways are similar in the initial step. After a double-strand break occurs, 5′ ends of the break are cut back to create 3′ overhangs of single-strand DNA (ssDNA). Recombinases bind the 3′ overhangs of ssDNA and search through vast quantities of DNA sequence to align and pair ssDNA with a homologous double-strand DNA template, facilitating the formation of a d-loop (Dunderdale et al., 1991; Murayama et al., 2008). In the DSBR pathway, the end of the invading 3′ strand is extended by a DNA polymerase and converted into a HJ. The other 3′ overhang strand also forms a HJ with the homologous DNA. After that, there are two pathways to convert the double HJs into recombination products (Sung and Klein, 2006). One pathway is mediated by HJ resolvases, which cleave HJs and produce either crossover or noncrossover products (Szostak et al., 1983). Various HJ resolvases have been found in archaea, bacteria, and eukaryotes (West, 2009). The eukaryotic nucleus also possesses another pathway, which is driven by the BTR complex consisting of the Bloom syndrome helicase, topoisomerase IIIα (TOP3A), and recombination-deficient Q-mediated genome instability subcomplex proteins (Wu and Hickson, 2003). The BTR complex does not cleave but dissolves the double HJs (Wu and Hickson, 2003). During the dissolution, the two HJ branches migrate toward each other until they form a hemicatenated intermediate, which is decatenated by TOP3A. Therefore, the dissolution pathway never produces crossover products (Supplemental Fig. S1). In contrast to DSBR, in the SDSA pathway, the invading 3′ ssDNA is utilized as a primer and extended along the recipient DNA duplex by a DNA polymerase without forming HJs. The newly synthesized strand dissociates from the template DNA and anneals with the other 3′ overhang strand through complementary base pairing. After the strands anneal, the remaining single-stranded gaps are filled by a DNA polymerase (Supplemental Fig. S1; Pâques and Haber, 1999).In nuclear, mitochondrial, and plastid genomes, numerous short dispersed repeats (SDRs) exist (Ottaviani et al., 2014). Thanks to the accuracy of recombinases in finding the homologous sequences despite the existence of myriad SDRs, the genomic sequence of the broken DNA is restored precisely via HR (McEntee et al., 1979; Shibata et al., 1979; Qi et al., 2015). However, in the absence of recombinases, SDRs anneal through complementary base pairing and produce recombinants between SDRs (Supplemental Fig. S1). Intramolecular recombination between SDRs results in deletion or inversion of the flanking region when they are oriented as direct or inverted repeats, respectively (Supplemental Fig. S2). This pathway is known as microhomology-mediated end joining (MMEJ; Supplemental Fig. S1; McVey and Lee, 2008).Recombinase genes, including Rad51 (eukaryotic type), radA (archaeal type), and recA (bacterial type), are believed to have evolved from a common ancestral gene (Lin et al., 2006). Two phylogenetically distinct RECA proteins are encoded in the nuclear genome of land plants, of which one is targeted to plastids (ptRECA) and the other to mitochondria (mtRECA; Lin et al., 2006). ptRECA and mtRECA are most closely related to cyanobacterial and proteobacterial counterparts, respectively, suggesting endosymbiotic origins of these proteins (Lin et al., 2006). Two observations should be made regarding the functions of these two plant recombinases: (1) Suppression or loss of function of ptRECA causes ectopic recombination between SDRs at a high frequency and destabilizes the plastid genome in the green alga C. reinhardtii and the moss Physcomitrella patens (Odahara et al., 2015a, 2016), suggesting that ptRECA maintains integrity of the plastid genome by promoting HR and thus suppressing MMEJ. Because MOC1 possesses HJ-resolving activity in vitro and is required for segregation of ptDNA in vivo (Kobayashi et al., 2017), RECA-mediated HR is likely accomplished with MOC1 through the DSBR pathway in plastids. However, how MOC1 functions in vivo has not been investigated. (2) Like ptRECA, mtRECA suppresses ectopic recombination between SDRs in the mitochondrial genome in P. patens and the seed plant Arabidopsis (Shedge et al., 2007; Odahara et al., 2009; Miller-Messmer et al., 2012). These results suggest that HJs are formed in plant mitochondria. However, except for CCE1, which is specific to yeasts (Saccharomyces cerevisiae and Schizosaccharomyces pombe; Kleff et al., 1992; Oram et al., 1998), mitochondrial HJ resolvases have not been identified in eukaryotes. Thus, it remains unclear whether HJs are formed and, if they are formed, how they are removed in mitochondria in plants.Regarding issues 1 and 2 described above, in this study, we show first that MOC1 suppresses ectopic recombination between SDRs in C. reinhardtii plastids, as does ptRECA, suggesting that MOC1 is involved in DSBR in the plastid. Next, we show that MOC1 dually targets plastids and mitochondria in the moss P. patens and maintains the integrity of ptDNA and mtDNA via suppression of ectopic recombination in both of these organelles. Putative dual-targeted transit peptides are also predicted in MOC1s of charophyte algae, a liverwort, a fern, and some seed plants, and we show that some of them are targeted to both plastids and mitochondria. Thus, it is suggested that HJs are formed during HR and removed by MOC1 in both plastids and mitochondria in some algae and plants.  相似文献   

13.
The L1CAM family of cell adhesion molecules is a conserved set of single-pass transmembrane proteins that play diverse roles required for proper nervous system development and function. Mutations in L1CAMs can cause the neurological L1 syndrome and are associated with autism and neuropsychiatric disorders. L1CAM expression in the mature nervous system suggests additional functions besides the well-characterized developmental roles. In this study, we demonstrate that the gene encoding the Caenorhabditis elegans L1CAM, sax-7, genetically interacts with gtl-2, as well as with unc-13 and rab-3, genes that function in neurotransmission. These sax-7 genetic interactions result in synthetic phenotypes that are consistent with abnormal synaptic function. Using an inducible sax-7 expression system and pharmacological reagents that interfere with cholinergic transmission, we uncovered a previously uncharacterized nondevelopmental role for sax-7 that impinges on synaptic function.  相似文献   

14.
The par genes of Caenorhabditis elegans are essential for establishment and maintenance of early embryo polarity and their homologs in other organisms are crucial polarity regulators in diverse cell types. Forward genetic screens and simple RNAi depletion screens have identified additional conserved regulators of polarity in C. elegans; genes with redundant functions, however, will be missed by these approaches. To identify such genes, we have performed a genome-wide RNAi screen for enhancers of lethality in conditional par-1 and par-4 mutants. We have identified 18 genes for which depletion is synthetically lethal with par-1 or par-4, or both, but produces little embryo lethality in wild type. Fifteen of the 18 genes identified in our screen are not previously known to function in C. elegans embryo polarity and 11 of them also increase lethality in a par-2 mutant. Among the strongest synthetic lethal genes, polarity defects are more apparent in par-2 early embryos than in par-1 or par-4, except for strd-1(RNAi), which enhances early polarity phenotypes in all three mutants. One strong enhancer of par-1 and par-2 lethality, F25B5.2, corresponds to nop-1, a regulator of actomyosin contractility for which the molecular identity was previously unknown. Other putative polarity enhancers identified in our screen encode cytoskeletal and membrane proteins, kinases, chaperones, and sumoylation and deubiquitylation proteins. Further studies of these genes should give mechanistic insight into pathways regulating establishment and maintenance of cell polarity.  相似文献   

15.
16.
Progress in understanding the molecular mechanism of self-assembly of amyloidogenic proteins and peptides requires knowledge about their structure in misfolded states. Structural studies of amyloid aggregates formed during the early aggregation stage are very limited. Atomic force microscopy (AFM) spectroscopy is widely used to analyze misfolded proteins and peptides, but the structural characterization of transiently formed misfolded dimers is limited by the lack of computational approaches that allow direct comparison with AFM experiments. Steered molecular dynamics (SMD) simulation is capable of modeling force spectroscopy experiments, but the modeling requires pulling rates 107 times higher than those used in AFM experiments. In this study, we describe a computational all-atom Monte Carlo pulling (MCP) approach that enables us to model results at pulling rates comparable to those used in AFM pulling experiments. We tested the approach by modeling pulling experimental data for I91 from titin I-band (PDB ID: 1TIT) and ubiquitin (PDB ID: 1UBQ). We then used MCP to analyze AFM spectroscopy experiments that probed the interaction of the peptides [Q6C] Sup35 (6–13) and [H13C] Aβ (13–23). A comparison of experimental results with the computational data for the Sup35 dimer with out-of-register and in-register arrangements of β-sheets suggests that Sup35 monomers adopt an out-of-register arrangement in the dimer. A similar analysis performed for Aβ peptide demonstrates that the out-of-register antiparallel β-sheet arrangement of monomers also occurs in this peptide. Although the rupture of hydrogen bonds is the major contributor to dimer dissociation, the aromatic-aromatic interaction also contributes to the dimer rupture process.  相似文献   

17.
Although regulatory small RNAs have been reported in photosynthetic cyanobacteria, the lack of clear RNA chaperones involved in their regulation poses a conundrum. Here, we analyzed the full complement of cellular RNAs and proteins using gradient profiling by sequencing (Grad-seq) in Synechocystis 6803. Complexes with overlapping subunits such as the CpcG1-type versus the CpcL-type phycobilisomes or the PsaK1 versus PsaK2 photosystem I pre(complexes) could be distinguished, supporting the high quality of this approach. Clustering of the in-gradient distribution profiles followed by several additional criteria yielded a short list of potential RNA chaperones that include an YlxR homolog and a cyanobacterial homolog of the KhpA/B complex. The data suggest previously undetected complexes between accessory proteins and CRISPR-Cas systems, such as a Csx1-Csm6 ribonucleolytic defense complex. Moreover, the exclusive association of either RpoZ or 6S RNA with the core RNA polymerase complex and the existence of a reservoir of inactive sigma–antisigma complexes is suggested. The Synechocystis Grad-seq resource is available online at https://sunshine.biologie.uni-freiburg.de/GradSeqExplorer/ providing a comprehensive resource for the functional assignment of RNA–protein complexes and multisubunit protein complexes in a photosynthetic organism.

We analyze a cyanobacterium using Grad-seq, providing a comprehensive resource for the in-depth analysis of the complexome in a photosynthetic organism.  相似文献   

18.
19.
G-quadruplex DNA structures have become attractive drug targets, and native mass spectrometry can provide detailed characterization of drug binding stoichiometry and affinity, potentially at high throughput. However, the G-quadruplex DNA polymorphism poses problems for interpreting ligand screening assays. In order to establish standardized MS-based screening assays, we studied 28 sequences with documented NMR structures in (usually ∼100 mM) potassium, and report here their circular dichroism (CD), melting temperature (Tm), NMR spectra and electrospray mass spectra in 1 mM KCl/100 mM trimethylammonium acetate. Based on these results, we make a short-list of sequences that adopt the same structure in the MS assay as reported by NMR, and provide recommendations on using them for MS-based assays. We also built an R-based open-source application to build and consult a database, wherein further sequences can be incorporated in the future. The application handles automatically most of the data processing, and allows generating custom figures and reports. The database is included in the g4dbr package (https://github.com/EricLarG4/g4dbr) and can be explored online (https://ericlarg4.github.io/G4_database.html).  相似文献   

20.
The lifestyle of parasitic plants is associated with peculiar morphological, genetic, and physiological adaptations that existing online plant-specific resources fail to adequately represent. Here, we introduce the Web Application for the Research of Parasitic Plants (WARPP) as an online resource dedicated to advancing research and development of parasitic plant biology. WARPP is a framework to facilitate international efforts by providing a central hub of curated evolutionary, ecological, and genetic data. The first version of WARPP provides a community hub for researchers to test this web application, for which curated data revolving around the economically important Broomrape family (Orobanchaceae) is readily accessible. The initial set of WARPP online tools includes a genome browser that centralizes genomic information for sequenced parasitic plant genomes, an orthogroup summary detailing the presence and absence of orthologous genes in parasites compared with nonparasitic plants, and an ancestral trait explorer showing the evolution of life-history preferences along phylogenies. WARPP represents a project under active development and relies on the scientific community to populate the web app’s database and further the development of new analysis tools. The first version of WARPP can be securely accessed at https://parasiticplants.app. The source code is licensed under GNU GPLv2 and is available at https://github.com/wickeLab/WARPP.

The WARPP online resource is a new, expandable, and interactive parasitic plant-specific data hub that provides online tools tailored to the peculiarities of parasitic angiosperms.  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