13 research outputs found

    Stationary phase expression of the arginine biosynthetic operon argCBH in Escherichia coli

    Get PDF
    BACKGROUND: Arginine biosynthesis in Escherichia coli is elevated in response to nutrient limitation, stress or arginine restriction. Though control of the pathway in response to arginine limitation is largely modulated by the ArgR repressor, other factors may be involved in increased stationary phase and stress expression. RESULTS: In this study, we report that expression of the argCBH operon is induced in stationary phase cultures and is reduced in strains possessing a mutation in rpoS, which encodes an alternative sigma factor. Using strains carrying defined argR, and rpoS mutations, we evaluated the relative contributions of these two regulators to the expression of argH using operon-lacZ fusions. While ArgR was the main factor responsible for modulating expression of argCBH, RpoS was also required for full expression of this biosynthetic operon at low arginine concentrations (below 60 μM L-arginine), a level at which growth of an arginine auxotroph was limited by arginine. When the argCBH operon was fully de-repressed (arginine limited), levels of expression were only one third of those observed in ΔargR mutants, indicating that the argCBH operon is partially repressed by ArgR even in the absence of arginine. In addition, argCBH expression was 30-fold higher in ΔargR mutants relative to levels found in wild type, fully-repressed strains, and this expression was independent of RpoS. CONCLUSION: The results of this study indicate that both derepression and positive control by RpoS are required for full control of arginine biosynthesis in stationary phase cultures of E. coli

    Alternative splicing of coq-2 controls the levels of rhodoquinone in animals

    Get PDF
    Parasitic helminths use two benzoquinones as electron carriers in the electron transport chain. In normoxia, they use ubiquinone (UQ), but in anaerobic conditions inside the host, they require rhodoquinone (RQ) and greatly increase RQ levels. We previously showed the switch from UQ to RQ synthesis is driven by a change of substrates by the polyprenyltransferase COQ-2 (Del Borrello et al., 2019; Roberts Buceta et al., 2019); however, the mechanism of substrate selection is not known. Here, we show helminths synthesize two coq-2 splice forms, coq-2a and coq-2e, and the coq-2e-specific exon is only found in species that synthesize RQ. We show that in Caenorhabditis elegans COQ-2e is required for efficient RQ synthesis and survival in cyanide. Importantly, parasites switch from COQ-2a to COQ-2e as they transit into anaerobic environments. We conclude helminths switch from UQ to RQ synthesis principally via changes in the alternative splicing of coq-2.Agencia Nacional de Investigación e InnovaciónCanadian Institutes of Health Researc

    Alternative splicing of coq-2 controls the levels of rhodoquinone in animals

    Get PDF
    Parasitic helminths use two benzoquinones as electron carriers in the electron transport chain. In normoxia, they use ubiquinone (UQ), but in anaerobic conditions inside the host, they require rhodoquinone (RQ) and greatly increase RQ levels. We previously showed the switch from UQ to RQ synthesis is driven by a change of substrates by the polyprenyltransferase COQ-2 (Del Borrello et al., 2019; Roberts Buceta et al., 2019); however, the mechanism of substrate selection is not known. Here, we show helminths synthesize two coq-2 splice forms, coq-2a and coq-2e, and the coq-2e-specific exon is only found in species that synthesize RQ. We show that in Caenorhabditis elegans COQ-2e is required for efficient RQ synthesis and survival in cyanide. Importantly, parasites switch from COQ-2a to COQ-2e as they transit into anaerobic environments. We conclude helminths switch from UQ to RQ synthesis principally via changes in the alternative splicing of coq-2.Agencia Nacional de Investigación e InnovaciónCanadian Institutes of Health Researc

    Caenorhabditis elegans is a useful model for anthelmintic discovery

    Get PDF
    Parasitic nematodes infect one quarter of the world's population and impact all humans through widespread infection of crops and livestock. Resistance to current anthelmintics has prompted the search for new drugs. Traditional screens that rely on parasitic worms are costly and labour intensive and target-based approaches have failed to yield novel anthelmintics. Here, we present our screen of 67,012 compounds to identify those that kill the non-parasitic nematode Caenorhabditis elegans. We then rescreen our hits in two parasitic nematode species and two vertebrate models (HEK293 cells and zebrafish), and identify 30 structurally distinct anthelmintic lead molecules. Genetic screens of 19 million C. elegans mutants reveal those nematicides for which the generation of resistance is and is not likely. We identify the target of one lead with nematode specificity and nanomolar potency as complex II of the electron transport chain. This work establishes C. elegans as an effective and cost-efficient model system for anthelmintic discovery

    An intestinally secreted host factor promotes microsporidia invasion of C. elegans.

    No full text
    Microsporidia are ubiquitous obligate intracellular pathogens of animals. These parasites often infect hosts through an oral route, but little is known about the function of host intestinal proteins that facilitate microsporidia invasion. To identify such factors necessary for infection by Nematocida parisii, a natural microsporidian pathogen of Caenorhabditis elegans, we performed a forward genetic screen to identify mutant animals that have a Fitness Advantage with Nematocida (Fawn). We isolated four fawn mutants that are resistant to Nematocida infection and contain mutations in T14E8.4, which we renamed aaim-1 (Antibacterial and Aids invasion by Microsporidia). Expression of AAIM-1 in the intestine of aaim-1 animals restores N. parisii infectivity and this rescue of infectivity is dependent upon AAIM-1 secretion. N. parisii spores in aaim-1 animals are improperly oriented in the intestinal lumen, leading to reduced levels of parasite invasion. Conversely, aaim-1 mutants display both increased colonization and susceptibility to the bacterial pathogen Pseudomonas aeruginosa and overexpression ofaaim-1 reduces P. aeruginosa colonization. Competitive fitness assays show that aaim-1 mutants are favored in the presence of N. parisii but disadvantaged on P. aeruginosa compared to wild-type animals. Together, this work demonstrates how microsporidia exploits a secreted protein to promote host invasion. Our results also suggest evolutionary trade-offs may exist to optimizing host defense against multiple classes of pathogens

    The novel nematicide wact-86 interacts with aldicarb to kill nematodes

    No full text
    <div><p>Parasitic nematodes negatively impact human and animal health worldwide. The market withdrawal of nematicidal agents due to unfavourable toxicities has limited the available treatment options. In principle, co-administering nematicides at lower doses along with molecules that potentiate their activity could mitigate adverse toxicities without compromising efficacy. Here, we screened for new small molecules that interact with aldicarb, which is a highly effective treatment for plant-parasitic nematodes whose toxicity hampers its utility. From our collection of 638 worm-bioactive compounds, we identified 20 molecules that interact positively with aldicarb to either kill or arrest the growth of the model nematode <i>Caenorhabditis elegans</i>. We investigated the mechanism of interaction between aldicarb and one of these novel nematicides called wact-86. We found that the carboxylesterase enzyme GES-1 hydrolyzes wact-86, and that the interaction is manifested by aldicarb’s inhibition of wact-86’s metabolism by GES-1. This work demonstrates the utility of <i>C</i>. <i>elegans</i> as a platform to search for new molecules that can positively interact with industrial nematicides, and provides proof-of-concept for prospective discovery efforts.</p></div

    Expression of in WT, Δ, Δand ΔΔstrains on LB plates containing X-Gal

    No full text
    <p><b>Copyright information:</b></p><p>Taken from "Stationary phase expression of the arginine biosynthetic operon in "</p><p>BMC Microbiology 2006;6():14-14.</p><p>Published online 22 Feb 2006</p><p>PMCID:PMC1413537.</p><p>Copyright © 2006 Weerasinghe et al; licensee BioMed Central Ltd.</p

    Wact-86 resistant mutants harbour missense mutations in the carboxylesterase gene <i>ges-1</i>.

    No full text
    <p><b>(A)</b> wact-86 dose-response assays for wild-type worms and the three wact-86 resistant mutants. For each resistant strain the <i>ges-1</i> allele and the GES-1 amino acid substitution are indicated (see <a href="http://www.plosntds.org/article/info:doi/10.1371/journal.pntd.0005502#pntd.0005502.s010" target="_blank">S2 File</a> for the whole genome sequencing data obtained for the resistant mutants). <b>(B)</b> Sequence alignment of the <i>C</i>. <i>elegans</i> GES-1 protein with the orthologous carboxylesterases from fly, fish, mouse, and human. For clarity, only the segment that is mutated in the wact-86 resistant strains is shown. Conserved residues are highlighted in black. The two GES-1 residues that are mutated in the wact-86 resistant strains are highlighted in grey. The asterisk denotes the conserved histidine that is part of the enzyme’s catalytic triad.</p

    The novel nematicide wact-86 interacts with aldicarb to kill <i>C</i>. <i>elegans</i>.

    No full text
    <p><b>(A)</b> The chemical structure of wact-86. <b>(B)</b> Combination dose-response matrix for wact-86 and aldicarb. Worm abundance, relative to the DMSO control, is represented by a colour-coded scale ranging from 0 (no viable worms) to ≥1 (at least as many viable worms as DMSO control). See <a href="http://www.plosntds.org/article/info:doi/10.1371/journal.pntd.0005502#sec009" target="_blank">Methods</a> for how the relative worm abundance value was calculated.</p
    corecore