crossmark THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 291, NO. 22, pp. 11928 –11938, May 27, 2016 © 2016 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A.

Role of Glyoxylate Shunt in Oxidative Stress Response*□ S

Received for publication, December 2, 2015, and in revised form, March 23, 2016 Published, JBC Papers in Press, April 1, 2016, DOI 10.1074/jbc.M115.708149

Sungeun Ahn‡, Jaejoon Jung‡, In-Ae Jang‡, Eugene L. Madsen§, and Woojun Park‡1 From the ‡Laboratory of Molecular Environmental Microbiology, Department of Environmental Science and Ecological Engineering, Korea University, Seoul 02841, Republic of Korea and the §Department of Microbiology, Cornell University, Ithaca, New York 14853 The glyoxylate shunt (GS) is a two-step metabolic pathway (isocitrate lyase, aceA; and malate synthase, glcB) that serves as an alternative to the tricarboxylic acid cycle. The GS bypasses the carbon dioxide-producing steps of the tricarboxylic acid cycle and is essential for acetate and fatty acid metabolism in bacteria. GS can be up-regulated under conditions of oxidative stress, antibiotic stress, and host infection, which implies that it plays important but poorly explored roles in stress defense and pathogenesis. In many bacterial species, including Pseudomonas aeruginosa, aceA and glcB are not in an operon, unlike in Escherichia coli. In P. aeruginosa, we explored relationships between GS genes and growth, transcription profiles, and biofilm formation. Contrary to our expectations, deletion of aceA in P. aeruginosa improved cell growth under conditions of oxidative and antibiotic stress. Transcriptome data suggested that aceA mutants underwent a metabolic shift toward aerobic denitrification; this was supported by additional evidence, including up-regulation of denitrification-related genes, decreased oxygen consumption without lowering ATP yield, increased production of denitrification intermediates (NO and N2O), and increased cyanide resistance. The aceA mutants also produced a thicker exopolysaccharide layer; that is, a phenotype consistent with aerobic denitrification. A bioinformatic survey across known bacterial genomes showed that only microorganisms capable of aerobic metabolism possess the glyoxylate shunt. This trend is consistent with the hypothesis that the GS plays a previously unrecognized role in allowing bacteria to tolerate oxidative stress.

The glyoxylate shunt (GS)2 is known to be essential for utilizing acetate and fatty acids as carbon sources under physiological conditions requiring gluconeogenesis (1). The glyoxylate cycle is a variant of the tricarboxylic acid cycle and shares five of the eight enzymes; the glyoxylate cycle bypasses the carbon dioxide-generating steps of the tricarboxylic acid cycle,

* This work was supported by a National Research Foundation of Korea (NRF) grant supported by the Korea government (MSIP) (NRF2014R1A2A2A05007010). The authors declare that they have no conflicts of interest with the contents of this article. □ S This article contains supplemental Tables S1–S5. The microarray data were deposited in the NCBI GEO site under accession no. GSE78230. 1 To whom correspondence should be addressed: Dept. of Environmental Science and Ecological Engineering, Korea University, Anam-Dong, Seungbuk-Ku, Seoul, 02841, Republic of Korea. Tel.: 82-2-3290-3067; Fax: 82-2953-0737; E-mail: [email protected] 2 The abbreviations used are: GS, glyoxylate shunt; PQ, paraquat; EPS, exopolysaccharide; ROS, reactive oxygen species; qRT, quantitative real-time.

11928 JOURNAL OF BIOLOGICAL CHEMISTRY

which are catalyzed by isocitrate dehydrogenase and ␣-ketoglutarate dehydrogenase (2). This shortcut for the tricarboxylic acid cycle has the potential to disturb cellular redox potentials because the primary function of the tricarboxylic acid cycle is the generation of reduced electron carriers (e.g. NADH and FADH2) via oxidation of acetyl CoA. Bypassing the complete tricarboxylic acid cycle conserves carbon atoms for gluconeogenesis while simultaneously diminishing the flux of electrons funneled into respiration (3). The product of aceA, the first enzyme of the glyoxylate shunt, catalyzes the formation of glyoxylate and succinate from isocitrate; the next reaction, conversion of glyoxylate to malate, proceeds via malate synthase, which is encoded by glcB (or aceB in Escherichia coli). Isocitrate lyase competes with isocitrate dehydrogenase because they share the same substrate, isocitrate. Their relative activities are controlled by isocitrate dehydrogenase kinase/phosphatase (AceK) through isocitrate dehydrogenase inactivation (4). The glyoxylate shunt is known to be up-regulated when acetyl-CoA is a direct product of a metabolic pathway, for example via degradation of acetate, fatty acids, and alkanes (5). However, a growing body of evidence suggests that the glyoxylate shunt also plays an important role in pathogenesis and the response to oxidative stress (6). Our research group has shown that hexadecane degradation by the oil-degrading bacterium Acinetobacter oleivorans DR1 generates oxidative stress, which induces expression of aceA, encoding an isocitrate lyase (7). Pectin degradation by Alishewanella spp. is also accompanied by oxidative stress, which triggers induction of the glyoxylate shunt (8). It seems likely that oxidative stress is an inevitable consequence of aerobic hydrocarbon degradation when there is a high demand for oxygen and possible uncoupling of oxygenase activity and in which side reactions can generate hydrogen peroxide (9). Bactericidal antibiotics commonly induce oxidative stress, which achieves cell death by promoting the formation of hydroxyl radicals through the Fenton reaction (10). Hydroxyl radical-mediated oxidative stress can be toxic by damaging proteins, membranes, and DNA (11). Mycobacterium tuberculosis was found to up-regulate its glyoxylate shunt when exposed to three different kinds of clinically used tuberculosis antibiotics: isoniazid, rifampicin, and streptomycin (12). E. coli experiencing superoxide stress also increased metabolic fluxes (48%) through the glyoxylate shunt (13). However, the molecular mechanism underlying the link between oxidative stress and the glyoxylate shunt remains poorly understood. Given its potential roles in oxidative stress defense, antibiotic resistance, and pathogenesis, the glyoxylate shunt is a promising target for development of new antibiotics. Furthermore, the VOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response glyoxylate shunt is widely distributed in bacteria, fungi, some protists, and plants (14) but is absent in most vertebrates. Bacterial responses to oxidative stress and the accompanying bacterial physiological changes have been extensively investigated in many pathogenic microorganisms, including Pseudomonas aeruginosa, which invades the lungs of cystic fibrosis patients and causes a potentially lethal infection (15). Oxidative stress causes many phenotypic and physiological changes in P. aeruginosa, including changes in biofilm production and pathogenesis (16, 17). P. aeruginosa can utilize acetate and fatty acids as sole carbon sources through the glyoxylate pathway. In P. aeruginosa, aceA (PA2634) is known to be important for assimilation of acyclic terpenes, leucine, ethanol, and acetate (18). The aim of this study was to explore a new role for the glyoxylate shunt in contributing to tolerance to oxidative stress. We applied microarray, biochemical, and physiological analyses to wild-type and both aceA and glcB knock-out mutants of P. aeruginosa. Furthermore, a global bioinformatics analysis across sequenced bacterial genomes suggests that this role may extend broadly across the bacterial domain.

Experimental Procedures Antibiotics, Culture Media, Bacterial Strains, and Growth Conditions—The wild-type and isocitrate lyase (aceA) and malate synthase (glcB) mutant strains of P. aeruginosa MPAO1 were purchased from the Washington University Genome Center. Bacteria were grown at 37 °C in Luria-Bertani (LB) broth or M9 minimal medium with glucose (2 g/liter), shaking at 220 rpm. M9 media were prepared with different sole carbon sources: glucose (0.2%), succinate (10 mM), acetate (10 mM), and hexadecanoic acid (0.2%). Paraquat (PQ) was added at the final concentrations described in the figure legends. Growth was monitored by measuring the A600 of the liquid cultures by using a BioPhotometer (Eppendorf). Gene Expression Analysis Using Real-time RT-PCR (qRTPCR)—Wild-type and mutant cells were grown to the exponential phase at 37 °C in LB. The cells were treated with 1 mM PQ and then incubated for 15 min. Isolation of total RNA, cDNA synthesis, and PCR were performed as previously described (19). The qRT-PCR was performed using the iCycler iQ realtime PCR detection system (Bio-Rad). For quantification, the 16S rRNA gene was used to obtain reference expression data. Three independent PCR reactions from two independent biological replicates were performed to measure relative expression of denitrification genes and oxidative stress-related genes. Primers used in qRT-PCR assay were listed in supplemental Table S1. Survival and Sensitivity of P. aeruginosa—For the sensitivity assay using paper disks, stationary-phase cells were added to each M9 agar plate with 0.2% glucose, then 20 ␮l each of 10 mM or 100 mM kanamycin was loaded into the plates on paper disks. After 24 h of incubation, bacteria-free zones around the disks were measured. For the survival test, 2.5 mM PQ or 3 mM KCN was used. For the PQ survival test, exponential-phase cells were harvested and washed 3 times with PBS (pH 7.4). Approximately 107 cell/ml were inoculated into fresh PBS (2 ml) containing PQ. For the KCN survival test, 3 mM aliquots were MAY 27, 2016 • VOLUME 291 • NUMBER 22

added with or without 1 mM sodium nitrate to LB. At each time point the cells were harvested and washed in PBS. The number of viable cells was determined by measuring the cfu on LB plates. Microarray Analysis—Cells were grown to the exponential phase (A600 ⫽ 0.4) at 37 °C with aeration. Cells were then treated with 1 mM PQ for 15 min. Total RNA was isolated using the RNeasy Mini kit (Qiagen). Specific steps for further RNA isolation were conducted as previously described (20). cDNA probes for cDNA microarray analysis were prepared by the reverse transcription of total RNA (50 ␮g) in the presence of aminoallyl-dUTP and 6 ␮g of random primers (Invitrogen) for 3 h. The cDNA probes were cleaned up using a Microcon YM-30 column (Millipore) followed by coupling to the Cy3 dye (for the reference) or Cy5 dye (for the test samples) (Amersham Biosciences). The dried Cy3- or Cy5-labeled cDNA probes were then resuspended in hybridization buffer containing 30% formamide (v/v), 5⫻ saline-sodium citrate, 0.1% SDS (w/v), and 0.1 mg/ml salmon sperm DNA. The Cy3- or Cy5-labeled cDNA probes were mixed together and hybridized onto a microarray slide. The hybridization images on the slides were scanned using an Axon 4000B (Axon Instruments) and analyzed with GenePix Pro software (version 3.0, Axon Instruments) to determine gene expression ratios (control versus test sample). The microarray data were deposited in the NCBI GEO site (under accession no. GSE78230). Genes that showed changes of ⬎1.5fold (up-regulated genes) were selected. Expression of representative genes was also confirmed by qRT-PCR. Measurement of Oxygen Consumption Rate—Exponentialphase cells grown in LB were diluted in 20 ml of LB in serum vials (radius, 30 mm; height, 60-mm clear serum vials were used; the volume of all serum vials was 27 ml), and further inflow of oxygen was blocked. Under these conditions the consumption of dissolved oxygen in the sealed 27 ml was measured with a Fibox 3 LCD trace transmitter (PreSens). Measurements were taken every 5 s for 30 min and analyzed with LCDTRACEv204 software. The rate of oxygen consumption was calculated as the decrease in oxygen concentration measured at 10-min intervals. Measurements were adjusted according to cell numbers (cfu) present in each vial. Experiments were conducted in triplicate. Detection of NO—The concentration of NO produced was estimated using the reagent 4,5-diaminofluorescein diacetate((Sigma), according to a previously described method (21). Exponential phase cultures (1 ml) were obtained and 10 ␮M 4,5-diaminofluorescein diacetate was added. After 1 h of incubation, cells were washed twice with PBS, and fluorescence was measured under green fluorescent light (excitation wavelength, 495 nm; emission wavelength, 515 nm) using a multidetection microplate reader (Hidex). Detection of N2O Concentration—Cells were diluted in 10 ml of LB in serum vials, and inflow of oxygen was blocked using a 20-mm silicon stopper that trapped gases generated during growth. After incubation, head-space gas was sampled through the stopper using a gas-tight syringe (Hamilton) at each time interval. The concentration of N2O was measured with a gas chromatograph (GC-2010; Shimadzu; ZP-Porapak Q 80/100 mesh column was used with the electron capture detector JOURNAL OF BIOLOGICAL CHEMISTRY

11929

Role of Glyoxylate Shunt in Oxidative Stress Response (ECD). The result of 4 h of incubation was depicted on a graph because nitrous oxide concentrations in 1- and 2-h samples were not enough for detection. Biofilm Formation Assay—Polystyrene 96-well microtiter plates (BD Biosciences) were used as abiotic surfaces to study biofilm formation. Bacterial cultures were grown overnight, washed twice in PBS, and inoculated at concentrations of 106 cell/ml in LB broth. After 24 h of incubation at 37 °C, biofilm formation was determined via crystal violet staining and quantified by measuring the absorbance at 595 nm, normalized by the absorbance at 600 nm. Exopolysaccharide (EPS) Analysis—The EPS was isolated as follows. First, culture solutions of P. aeruginosa wild type (WT) and mutants grown in LB media were centrifuged for 30 min at 12,000 ⫻ g at 4 °C. Next, the supernatants were mixed with three volumes of cold ethanol and maintained overnight at 4 °C. After centrifugation (5000 ⫻ g, 15 min, 4 °C), the pellets were suspended in 80% ethanol and centrifuged again followed by three rounds of washing. The final precipitates were dissolved in distilled water. To remove excess salts, the EPS was dialyzed (Centricon, Amicon). Then the extracted EPS was freeze-dried and weighed. Confocal Laser Scanning Microscopy—Cells from biofilms were stained with 200 ␮l of SYPRO Ruby biofilm matrix stain (Invitrogen) for 20 min. The cells were then observed by confocal laser scanning microscopy (LSM700; Carl Zeiss). Before the biofilm staining procedure, cells were incubated on the cover glass for 12 h at 37 °C. Confocal images of SYPRO Rubystained biofilm cells were observed under red fluorescent light (excitation wavelength, 450 nm; emission wavelength, 610 nm) to evaluate the height and density of the biofilms (C-Apochromat 403/1.20 W Corr M27; Carl Zeiss). The confocal laser scanning microscopy images were analyzed using the Zen 2012 program. Determination of ATP Concentrations—To measure intracellular ATP concentrations, the ENLITEN ATP assay system bioluminescence detection kit (Promega) was used in accordance with the manufacturer’s instructions. To confirm ATP concentrations, exponentially growing cultures were treated with 0.5 mM or 1 mM PQ for 15 min, and cells were harvested for ATP extraction using 1% trichloroacetic acid buffer and treated as previously described (22). The ATP concentration is expressed as the molar concentration per mg of protein. Measurement of the NADH/NAD⫹ Ratio—Nicotinamide adenine dinucleotide (NAD⫹) and NADH concentrations were measured using the EnzyChromTM NAD⫹/NADH Assay kit (BioAssay Systems) according to the manufacturer’s instructions. Exponentially growing cultures were treated with 1 mM PQ for 15 min, and cells were collected for NADH and NAD⫹ extraction. Further procedures were performed as previously described (22). NAD⫹ and NADH concentrations were normalized by the amount of total protein. Determination of Cellular Acid-soluble Thiols—Total acidsoluble thiols were measured using the method of Lawley and Thatcher (57), with some modifications. WT, ⌬aceA, and ⌬glcB cells grown in 50 ml of M9 media to the exponential phase were centrifuged and washed in phosphate buffer with 1 mM EDTA. The cells were lysed by sonication. Proteins were precipitated

11930 JOURNAL OF BIOLOGICAL CHEMISTRY

from the supernatant fractions by the addition of cold trichloroacetic acid (final concentration of 5%). After 10 min the acidified samples were clarified by centrifugation, and 20 ␮l of clear supernatant was added to 180 ␮l of a 5,5⬘-dithiobis-(2-nitrobenzoic acid) (DTNB) solution (200 ␮g/ml in 0.2 M sodium phosphate buffer). The absorption at 410 nm was measured immediately. Glutathione (GSH) was used as the thiol standard, and the thiol content was normalized to the amount of total protein. Bioinformatic Analysis of the Glyoxylate Shunt—To investigate the phylogenetic distribution of the glyoxylate shunt, isocitrate lyase and malate synthase sequences were retrieved using the Gene Search function in the integrated microbial genome (IMG) database (23), which contained 27,319 bacterial genomes as of July 2015. To reduce database redundancy, only one strain was arbitrarily selected for each species when genomic data were available for multiple strains of one species. The oxygen requirements of selected bacterial species were identified based on experimental evidence from published papers. Bacterial species harboring the glyoxylate shunt and experimentally identified physiology were subjected to cladogram analysis. Cladograms in the Newick format were generated using the program phyloT and visualized using iTOL (24).

Results Physiological and Phenotypic Alterations in the Absence of the Glyoxylate Shunt—To ascertain the broad physiological role of the glyoxylate shunt in P. aeruginosa, PA2634 (⌬aceA) and PA0482 (⌬glcB) mutants were tested in a variety of assays. Based on the GS pathway stoichiometry, the aceA mutant is strongly committed to tricarboxylic acid cycle-related pathways that deliver reducing power (e.g. NADH and FADH2) to respiratory chains, whereas the glcB mutant likely experiences elevated levels of glyoxylate when the GS is induced. Growth of the two mutants was tested using minimal media containing different carbon sources (Fig. 1A). With succinate as sole carbon source, both mutants displayed growth curves similar to the WT. As expected, the aceA and glcB mutants showed growth defects in media requiring a functional glyoxylate shunt (acetate and hexanoic acid a carbon sources; Fig. 1A) (25). However, an unexpected growth promotion was observed in both mutants when grown in rich LB media that should boost oxidative stress, relative to the minimal medium. To further investigate this surprising growth-enhanced phenotype in the mutants, we raised the level of superoxide-induced oxidative stress in the cells by adding PQ to the LB medium; growth enhancement was retained in the aceA mutant, whereas the glcB mutant showed a severe growth defect (Fig. 1B). Note that a high intracellular glyoxylate concentration (expected in the glcB mutant) is likely to be toxic because glyoxylate can be transformed into toxic intermediates, such as glyoxal or glycolaldehyde (26). The contrasting responses of the two mutants to PQ (Fig. 1B) prompted us to measure the expression of both genes in WT cells exposed to PQ and hydrogen peroxide (H2O2); only aceA was highly up-regulated (Fig. 1C). Thus, aceA is clearly activated in response to oxidative stress. To confirm that our experimental procedures were valid, we performed a control experiVOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response

FIGURE 1. Characterization of the glyoxylate-shunt phenotype in P. aeruginosa using WT and mutant cells under a variety of conditions. A, influence of carbon sources on growth of WT and mutant cells of P. aeruginosa. Cellular growth curves were measured in minimal media containing either succinate, acetate, or hexanoic acid as sole carbon sources. Because of its low solubility, the growth in hexanoic acid was determined using cfu/ml. B, growth inhibition of WT, ⌬aceA, and ⌬glcB mutants in 0.5 mM PQ-supplemented LB. C, relative expression of aceA and glcB in WT P. aeruginosa under conditions of H2O2 and PQ treatment were analyzed using qRT-PCR. D, survival of WT, ⌬aceA, and ⌬glcB P. aeruginosa strains in fresh PBS supplemented with 2.5 mM of PQ. E, paper-disk assays show sensitivities of WT, ⌬aceA, and ⌬glcB strains of P. aeruginosa to kanamycin and PQ on M9 agar plate supplemented with 0.2% glucose. At least three independent experiments were performed in all experiments. Data represent the average of biological replicates with S.D. *, p ⬍ 0.05 relative to respective untreated control (⫺) or a value of 0 h by unpaired Student’s t test.

ment with an obligate aerobic strain of Pseudomonas putida, which showed that our system provided a truly anaerobic environment; without oxygen, aceA was not up-regulated in the presence of PQ (data not shown). A survival assay in broth culture dosed with PQ (2.5 mM) also showed that aceA mutant cells were more resistant to PQ toxicity than WT cells and the glcB mutant (Fig. 1D). Moreover, a paper disc assay with PQ and kanamycin confirmed the trends found for PQ survival assay in broth (Fig. 1E). The above observations clearly establish that aceA and glcB have distinctive physiological impacts on the cell physiology of P. aeruginosa. Transcriptomic and Biochemical Analyses of aceA Mutants— Microarray analyses of P. aeruginosa WT and glyoxylate shunt mutants were conducted with or without PQ. Up-regulation of genes involved in denitrification was immediately noticeable in mutants without PQ despite the fact that these assays were conducted under aerobic conditions (supplemental Table S2). MAY 27, 2016 • VOLUME 291 • NUMBER 22

Nitrate present in LB medium would likely readily be available as an electron acceptor for denitrification. The nitrate reductase operon (narGHJI), the nitrite reductase gene (nirS), the nitric oxide reductase operon (norDBC), and the nitrous-oxide reductase gene (nosZ), required for denitrification, were all highly up-regulated in aceA mutants, whereas a lower level of induction was observed in glcB mutants. qRT-PCR largely confirmed the microarray data (Table 1): especially in the case of the aceA mutant (Fig. 2A). Activation of nitrate-respiration genes implies alteration of cellular electron flow away from oxygen as the terminal electron acceptor. Oxygen consumption was measured simultaneously, and we found that the aceA mutant cells consumed less oxygen than WT cells (Fig. 2B), although the growth of aceA mutants was enhanced in LB medium (Fig. 1B). Potassium cyanide (KCN) is known to bind to cytochrome oxidase in cells, thereby interrupting oxygen transport and preventing aerobic respiration. The aceA JOURNAL OF BIOLOGICAL CHEMISTRY

11931

Role of Glyoxylate Shunt in Oxidative Stress Response TABLE 1 Expression analysis of denitrification genes using qRT-PCR in wild type and mutants of glyoxylate shunt gene

a

-Fold changea

Name

Locus tag

Protein name

⌬aceA/WT

⌬glcB/WT

narH napA narG norD norB narX nirS narJ nosZ norC narI napB

PA3874 PA1174 PA3875 PA0525 PA0524 PA3878 PA0519 PA3873 PA3392 PA0523 PA3872 PA1173

Respiratory nitrate reductase ␤ chain Periplasmic nitrate reductase protein NapA Respiratory nitrate reductase ␣ chain Probable denitrification protein NorD Nitric-oxide reductase subunit B Two-component sensor NarX Nitrite reductase precursor Respiratory nitrate reductase delta chain Nitrous-oxide reductase precursor Nitric-oxide reductase subunit C Respiratory nitrate reductase ␥ chain Cytochrome c-type protein NapB precursor

3.21 ⫾ 0.18 2.92 ⫾ 0.41 2.37 ⫾ 0.47 2.32 ⫾ 0.06 2.30 ⫾ 2.23 2.14 ⫾ 0.32 2.04 ⫾ 0.21 1.92 ⫾ 0.31 1.83 ⫾ 0.12 1.77 ⫾ 0.14 1.73 ⫾ 0.41 1.64 ⫾ 0.47

1.37 ⫾ 0.41 1.95 ⫾ 0.31 1.13 ⫾ 0.42 1.49 ⫾ 0.12 1.00 ⫾ 0.12 1.99 ⫾ 0.31 1.24 ⫾ 0.19 3.41 ⫾ 0.51 1.47 ⫾ 0.04 1.32 ⫾ 0.31 2.12 ⫾ 0.41 2.32 ⫾ 0.28

The average values with standard deviation were obtained from duplicate biological independent experiments, and qRT-PCR reaction was performed in triplicate. -Fold change represents relative expression to wild-type in exponential phase.

FIGURE 2. The isocitrate lyase (aceA)-deficient strain of P. aeruginosa exhibits respiration using nitrate. A, confirmation of expression of mRNA of denitrification-related genes by qRT-PCR. B, the O2 consumption rate was measured in WT, ⌬aceA, and ⌬glcB strains. Four independent experiments were performed. C, survival of WT, ⌬aceA, and ⌬glcB P. aeruginosa strains in LB medium supplemented with 3 mM KCN. D, survival of WT, ⌬aceA, and ⌬glcB P. aeruginosa strains in LB medium supplemented with KCN and 1 mM nitrate. E, production of nitric oxide by WT and ⌬aceA strains. The level of nitric oxide was measured using 4,5-diaminofluorescein diacetate fluorescence dye. F, production of nitrous oxide by WT, ⌬glcB, and ⌬aceA strains. The level of nitrous oxide was assessed using gas chromatography. ⌬aceA mutants showed increased levels of nitrous oxide. At least three independent experiments were performed in all experiments. Data represent the average of biological replicates with S.D. *, p ⬍ 0.05 relative to respective untreated control (⫺) or value of 0 h by unpaired Student’s t test.

11932 JOURNAL OF BIOLOGICAL CHEMISTRY

VOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response TABLE 2 Expression analysis of oxidative stress-related genes using qRT-PCR in wild type and mutants of glyoxylate shunt gene with or without 1 mM paraquat -Fold changea

a

Name

Locus tag

Protein name

⌬aceA/WT

⌬glcB/WT

WT PQ/WT

⌬aceA PQ/⌬aceA

⌬glcB PQ/⌬glcB

katN fpr sodB sodM ahpC trxA oxyR katA soxR katB katE

PA2185 PA3397 PA4366 PA4468 PA0139 PA5240 PA5344 PA4236 PA2273 PA4613 PA2147

Non-heme catalase KatN Ferredoxin-NADP⫹ reductase Superoxide dismutase Superoxide dismutase Alkyl hydroperoxide reductase subunit C Thioredoxin Transcriptional regulator Catalase Transcriptional regulator Catalase Catalase HPII

2.17 ⫾ 0.38 1.31 ⫾ 0.12 1.31 ⫾ 0.04 1.11 ⫾ 0.33 1.1 ⫾ 0.37 1.01 ⫾ 0.17 0.85 ⫾ 0.11 0.63 ⫾ 0.11 0.51 ⫾ 0.14 0.23 ⫾ 0.18 0.19 ⫾ 0.64

1.87 ⫾ 0.33 0.93 ⫾ 0.18 1.31 ⫾ 0.12 0.73 ⫾ 0.22 1.04 ⫾ 0.12 0.87 ⫾ 0.27 1.31 ⫾ 0.64 1.14 ⫾ 0.24 0.78 ⫾ 0.13 1.12 ⫾ 0.31 1.27 ⫾ 0.61

0.32 ⫾ 2.02 2.37 ⫾ 0.87 1.91 ⫾ 0.53 0.41 ⫾ 0.22 2.67 ⫾ 0.38 2.27 ⫾ 0.18 2.94 ⫾ 0.18 2.31 ⫾ 0.71 2.13 ⫾ 0.28 2.73 ⫾ 0.13 0.46 ⫾ 0.48

0.67 ⫾ 0.18 1.08 ⫾ 0.19 1.73 ⫾ 0.27 0.67 ⫾ 0.21 1.27 ⫾ 0.13 1.37 ⫾ 0.18 1.27 ⫾ 0.37 1.97 ⫾ 0.21 1.64 ⫾ 0.13 2.28 ⫾ 0.11 0.27 ⫾ 0.13

0.97 ⫾ 0.17 1.32 ⫾ 0.39 1.12 ⫾ 0.37 0.60 ⫾ 0.27 1.10 ⫾ 0.48 0.74 ⫾ 0.31 1.32 ⫾ 0.08 0.97 ⫾ 0.27 1.02 ⫾ 0.31 1.00 ⫾ 0.19 0.50 ⫾ 0.01

The average values with standard deviation were obtained from duplicate biological independent experiments, and qRT-PCR reaction was performed in triplicate. -Fold change represents relative expression to wild-type with or without 1 mM paraquat treatment.

mutants (featuring up-regulated denitrification genes; Table 1) would be expected to show enhanced cyanide resistance (Fig. 2C). Also, as predicted, when nitrate was added to the medium, both mutants showed greater resistance to KCN (Fig. 2D), suggesting that glcB mutants can also direct electron flow toward denitrification despite the likely intracellular accumulation of toxic glyoxylate. To confirm that up-regulated denitrification transcripts led to a respiratory shift toward denitrification, we quantified production of the denitrification intermediates, NO and N2O (Fig. 2, E and F). Genes encoding many membrane transporters were also highly up-regulated. Transporters in the major facilitator superfamily and the resistance-nodulation-division (RND) family showed the highest up-regulation in aceA mutants (supplemental Table S3). In particular, mexE (PA2493) showed a 54.454-fold increase in expression in aceA mutants compared with the WT. The up-regulation of this well known multidrug efflux pump system may also contribute to the tolerance of P. aeruginosa to toxic chemicals. Enhanced Biofilm Formation and Pigment Production in aceA Mutants—Our data showed that aceA mutants showed increased resistance to oxidants including PQ and kanamycin (Fig. 1E). The degree of gene expression involved in oxidative stress is less in PQ-treated mutants compared with PQ-treated wild-type cells. Interestingly, no significant expression of oxidative stress-related genes was observed in aceA or glcB mutants without PQ, except up-regulation of katN (PA2185), which encodes a non-heme catalase (Table 2, supplemental Table S4). In cell binding assays to microtiter plates, the aceA and glcB-deficient mutants showed increased biofilm formation (Fig. 3A). We hypothesize that the thick EPS layer may be beneficial for aerobic denitrification and resistance to chemical stressors. In fact, enhanced EPS production was observed in aceA mutants (Fig. 3B). Confocal microscopic analysis confirmed increased biofilm promotion in aceA mutants (Fig. 3C). Interestingly, when the mutant cells were cultivated, a distinct color change was observed, resulting from increased pyoverdin and pyocyanin formation (data not shown). Microarray analysis supported these findings (supplemental Table S3). Intracellular NADH Level with PQ Treatment —Exposure to PQ for 15 min increased ATP contents in all cells (Fig. 4A). This short term increase in ATP has been proposed to be caused by MAY 27, 2016 • VOLUME 291 • NUMBER 22

FIGURE 3. Enhanced intracellular ROS levels and biofilm formation in glyoxylate shunt-deficient strains. A, test of biofilm formation in WT, ⌬aceA, and ⌬glcB mutants. Three independent experiments were performed. B, quantification of EPS yields in WT and mutant strains. EPS formation was calibrated using cell weight, and results are expressed as yield. Two independent experiments were performed. C, a confocal image of the biofilm formed by each strain allows comparison of biofilm formation and shape. Data represent the average of biological replicates with S.D. *, p ⬍ 0.05 relative to respective untreated control (⫺) by unpaired Student’s t test. **, p ⬍ 0.05 ⌬glcB relative to ⌬aceA by unpaired Student’s t test.

inhibition of energy-consuming processes by PQ rather than increased respiration (27). However, PQ-exposed aceA mutants had a higher basal ATP concentration than the two other cell types, which might be caused by the growth condition that preceded the assay (Fig. 4A). Both WT and glcB mutants, which showed a growth defect upon PQ addition, displayed lower ATP concentrations. The intracellular NADH/NAD⫹ ratio might be affected by the availability of electron donors and acceptors (28). Both mutants had a lower basal NADH/NAD⫹ ratio because of their higher rates of respiration and efficient growth in LB (Fig. 4A). The NADH/NAD⫹ ratio under PQ conditions was lower in the WT, likely due to the fact that the glyoxylate shunt bypasses two NADH-generating steps. Cells show an increased NADH/NAD⫹ ratio when NADH oxidation is slowed by limitation of electron acceptors (29). The addition JOURNAL OF BIOLOGICAL CHEMISTRY

11933

Role of Glyoxylate Shunt in Oxidative Stress Response

FIGURE 4. Changes in redox states and glutathione levels under conditions of ROS in P. aeruginosa. A, the ATP, NADH/NAD⫹, and NADPH/NADP⫹ levels were indicators of redox states in WT and the aceA and glcB mutants. Redox levels were measured with or without 0.5 mM or 1 mM PQ. B, the total GSH level was measured in WT and mutant strains with and without 1 mM H2O2. At least three independent experiments were performed in all experiments. Data represent the average of biological replicates with S.D. *, p ⬍ 0.05 relative to respective untreated control (⫺) by unpaired Student’s t test.

of PQ did not alter the NADH/NAD⫹ ratio in the aceA mutants, indicating that respiration in these cells was maintained, perhaps through a combination of pathways; lower oxygen consumption and higher denitrification-related events were observed (Fig. 2A). The level of NADH in PQ-treated WT cells dropped markedly; a similar observation has been made after antibiotic treatment (29). Our data suggested that WT cells experienced increased oxidative stress when treated with PQ, whereas mutants did not. In the mutants only a subtle change in the NADPH/NADP⫹ ratio was observed after PQ addition (Fig. 4A). Total Thiol Levels under ROS Stress—Like other ␣-keto acids (30), glyoxylate can react with H2O2, thus providing an indirect defense against ROS stress. Glyoxylate may also play a role in ROS defense by generating glycine via alanine-glyoxylate transaminase or serine-glyoxylate transaminase, which could lead to the production of glutathione, a tripeptide antioxidant composed of glycine-cysteine-glutamate (31). Interestingly, transaminase (PA0132) was one of the most highly up-regulated genes in PQ-treated WT cells (supplemental Table S5). Expression analysis combined with total thiol quantification showed that the level of glutathione, which composed the majority (⬎95%) of total thiols, was higher in WT than the two mutants (Fig. 4B). The gene encoding glutathione synthase (PA0407) was up-regulated 1.845-fold in PQ-treated WT cells, whereas aceA and glcB mutants showed no significant changes (supplemental Table S5). Genes encoding glutathione S-transferase (PA1655, PA1033, and PA2813) and those encoding glutathione peroxidases (PA2826 and PA0838), all of which are well known to function in ROS protection (32), were also up-regu-

11934 JOURNAL OF BIOLOGICAL CHEMISTRY

lated under PQ in the WT (supplemental Table S5). However, H2O2 treatment increased the total thiol level in both WT and aceA mutants, suggesting that glutathione synthesis may proceed via an alternative pathway under such conditions (Fig. 4B). Phylogenetic Distribution of the Glyoxylate Shunt in Bacteria—We next investigated the extent to which the glyoxylate shunt co-occurs with genes encoding aerobic respiration in bacteria. We investigated the phylogenetic distribution of the glyoxylate shunt across all available bacterial genomes. A search for glyoxylate shunt in the integrated microbial genomes database showed that it is present in 1538 bacterial genomes from 339 genera belonging to 7 phyla. To bioinformatically verify the completeness of the glyoxylate shunt, we confirmed the co-occurrence of isocitrate lyase and malate synthase. To determine the relationship between the glyoxylate shunt and various physiological or ecological characteristics, experimental data from previously published papers were compared. In total, 957 species were found to possess the phenotype in question, and the genomes of those species were used to construct a cladogram (Fig. 5). Proteobacteria (65.2%), Actinobacteria (20.9%), and Firmicutes (9.6%) were the major phyla found to possess the glyoxylate shunt, showing that this bypass is not randomly distributed among bacterial species. We observed that all species harboring the glyoxylate shunt were aerobes or facultative anaerobes, with the exception of the anaerobic species Thiorhodospira sibirica; bacteria with halophilic, acidophilic, alkaliphilic, or thermophilic lifestyles and habitats were not found to possess the shunt (33). The copy number of isocitrate lyase and malate synthase in most bacterial genomes was one or two. Thirty of the genomes were found to harbor more than three VOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response

FIGURE 5. Phylogenetic distribution of the glyoxylate shunt in bacteria. This cladogram shows 957 bacterial species that possess both isocitrate lyase and malate synthase. The overlapping pie chart indicates the taxonomic placement (phylum and class) of each node and the proportion. Species names are shown at branch ends. The outer circle (green) shows the glyoxylate shunt copy number in each genome (higher and lower bars indicate 1 and 2 copies, respectively). The inner circle shows the respiration characteristics of each strain.

copies of the glyoxylate shunt; however, these appeared to be incorrect annotations. Thus, our bioinformatics analysis indicates that aerobic respiration and the glyoxylate shunt are linked. We hypothesize that protection from ROS is the mechanism of this linkage.

Discussion Only one isocitrate lyase, encoded by the PA2634 gene, is present in the genome of P. aeruginosa (18). Little is known about the contribution of the glyoxylate shunt to bacterial physiology, except that it plays roles in acetate and fatty acid metabolism. However, it has been reported that the glyoxylate shunt, particularly isocitrate lyase, is required for the pathogenesis of M. tuberculosis, and aceA mutants showed significantly attenuated survival in macrophages (34). Similar findings have been reported for other pathogenic bacteria, such as P. aeruginosa and Rhodococcus equi, as well as for fungal strains (35–37). In the case of R. equi, the authors speculated that aceA could be involved in metabolism of host membrane lipid-derived fatty acids, which are potential carbon sources in macrophages (37). However, a full understanding of the relevance of the glyoxylate shunt under various stresses remains unclear (9, 19, 38). The role and regulation of the glyoxylate shunt have mainly been studied in E. coli, in which the glyoxylate shunt is present in an operon, aceBAK (2). In P. aeruginosa, aceA (PA2634) and glcB (PA0482, malate synthase) are present at different genomic locations (18), and our global genomic analysis demonstrated that this is also the case in many other species of bacteria. In E. coli, IclR negatively regulates the aceBAK operon and MAY 27, 2016 • VOLUME 291 • NUMBER 22

appears to be directly regulated by glyoxylate and pyruvate, which have been shown to bind to the C-terminal domain of IclR (39, 40). IclR is also indirectly regulated by FadR, a fatty acid biosynthesis regulator (41). Furthermore, positive regulation by both FruR and integration host factor (IHF) has been reported (42, 43). The indirect regulation of aceA by RpoN has been documented in P. aeruginosa (44), but other mechanisms for regulation of the glyoxylate shunt have not been well studied. In particular, it remains unclear how ROS regulate this pathway. Our data show that aceA and glcB were not induced to the same degree in P. aeruginosa in the presence of added PQ (Fig. 1C). This different degree of up-regulation has also been shown in our previous studies (9, 19, 38). We observed that aceA and glcB mutants showed different responses to PQ (Fig. 1, B–E), which was supported by other research (45). It is worth noting that the growth defect of glcB mutants under PQ was restored by cysteine but not by other metabolites such as glutamate, glycine, glutathione, or glyoxylate (data not shown). Contrary to our expectations, aceA mutants of P. aeruginosa had a growth advantage in rich media (Fig. 1B), in which elevated levels of ROS stress can be generated, as shown by our previous findings.3 Growth in succinate-supplemented minimal medium was equivalent in all strains. Likewise, the observed resistance of aceA mutants to antibiotics and PQ could be due to decreased ROS stress under such conditions (Fig. 1, B–D). Here, we found that enhanced growth of the aceA

3

S. Ahn, J. Jung, I.-A. Jang, E. L. Madsen, and W. Park, submitted for publication.

JOURNAL OF BIOLOGICAL CHEMISTRY

11935

Role of Glyoxylate Shunt in Oxidative Stress Response mutant in biofilms was linked to both flow of respiration toward denitrification and enhanced production of EPS (Figs. 2 and 3). The importance of EPS production, biofilm formation, and generation of outer membrane proteins, such as OprF, has been shown for P. aeruginosa infection of the cystic fibrosis lung environment, which has microaerobic oxygen conditions (46, 47). Cells of aceA mutants showed similar characteristics to cystic fibrosis-adjusted cell features, indicating that aceA mutants had adapted to avoid ROS stress by using both respiration systems. Increased EPS production, which may aid in formation of conditions under which denitrification is possible, might have resulted from the ⬎2.5-fold up-regulation of alginate synthase (PA3544) observed in aceA mutants (supplemental Table S3). EPS could restrict the passage of toxic chemicals, including PQ. This is supported by the fact that alginate production and biofilm formation are controlled by oxygen availability in P. aeruginosa (48). Aerobic denitrification in continuous cultures of P. aeruginosa has been demonstrated, even at high concentrations of dissolved oxygen (49). Interestingly, a linkage between the mexEF-oprN (PA2493-PA2495) operon and the nitrate respiratory chain has also been reported, even under aerobic conditions (50). Our data suggest that they may also play a role in aerobic denitrification in aceA mutants (supplemental Table S3). Additionally, up-regulation of glutathione-related genes was observed in PQ-treated WT cells, which could play an important role in ROS stress defense (supplemental Table S4). Due to the differential expression of aceA and glcB observed under PQ-treated conditions, we hypothesized that the glyoxylate produced by isocitrate lyase could be used in other mechanisms (Fig. 1C). It has been reported that Chlamydomonas reinhardtii regulates isocitrate lyase by glutathionylation (51) and that a transaminase (PA0132), which was highly up-regulated in PQtreated WT cells (supplemental Table S5), can transform glyoxylate to glycine, a component of glutathione. Further experiments are warranted to investigate the role of glyoxylate in the regulation and production of glutathione in P. aeruginosa (Fig. 4B). Involvement of the glyoxylate shunt in denitrification was first identified in Thiobacillus versutus (52). Previous studies focused on the metabolic function of the glyoxylate shunt reported induction by specific carbon sources such as acetate. However, many of our experiments here were performed in rich media (LB) to exclude the influence of acetate metabolism. Interestingly, up-regulated expression of katN (PA2185), which encodes a non-heme catalase, was observed in mutants (Table 2); however, its role in P. aeruginosa is not fully understood (53). Increased siderophore production and decreased susceptibility to iron chelators were also observed in mutant cells (data not shown). These data indicate that iron requirement for WT and mutants would differ. It has been shown that inhibition of oxygen transfer under conditions of iron deprivation enhances secretion of iron chelators in P. aeruginosa (54, 55). Taken together, these iron-related features are important for denitrification respiratory-induced cell conditions. In Neisseria gonorrhoeae, the electron transfer chains of aerobic respiration and denitrification have been shown to be interwoven, allowing N. gonorrhoeae to co-express aerobic respiration and denitrification in microaerobic conditions (56).

11936 JOURNAL OF BIOLOGICAL CHEMISTRY

The exact proportion of respiration that is performed via denitrification in aceA mutants remains unclear; however, we hypothesize that P. aeruginosa aceA mutants achieve a balance between the two respiratory pathways. Our bioinformatics analysis revealed that only aerobes or facultative anaerobes employ the glyoxylate shunt. This result implies that isocitrate lyase may also play a role in regulation of respiration and ROS management in many other species of bacteria (Fig. 5). The glyoxylate shunt role in bacterial physiology has traditionally been associated with the need for gluconeogenesis, as induced by carbon-source limitation. Here, we have shown that the glyoxylate shunt in P. aeruginosa is induced by oxidative stress. This type of oxidative stress has implications for both colonization of surfaces (biofilm formation) and genomic evolution across the bacterial domain. We have also revealed several aspects of the mechanism by which the glyoxylate shunt governs a shift from aerobic to nitrate-based cellular respiration. Future research advancing a deeper understanding of the connections between oxidative stress response, the glyoxylate shunt, biofilm formation, cell pathogenicity, and development of microbial communities is warranted. Author Contributions—S. A. and W. P. designed the study. S. A., J. J., and I.-A. J. performed the experiments. S. A., J. J., and I.-A. J. collected the data. S. A. and J. J. performed the statistical analyses and wrote the first complete draft of the manuscript. E. L. M. and W. P. provided substantial modifications, and all authors contributed to the final version and approved it.

References 1. Maloy, S. R., Bohlander, M., and Nunn, W. D. (1980) Elevated levels of glyoxylate shunt enzymes in Escherichia coli strains constitutive for fatty acid degradation. J. Bacteriol. 143, 720 –725 2. Kornberg, H. L. (1966) The role and control of the glyoxylate cycle in Escherichia coli. Biochem. J. 99, 1–11 3. White, D., Drummond, J. and Fuqua, C. (2012) The Physiology and Biochemistry of Prokaryotes, 4th Ed., pp. 214 –237, Oxford, University Press, New York, NY 4. Ikeda, T., and LaPorte, D. C. (1991) Isocitrate dehydrogenase kinase/ phosphatase: aceK alleles that express kinase but not phosphatase activity. J. Bacteriol. 173, 1801–1806 5. Renilla, S., Bernal, V., Fuhrer, T., Castaño-Cerezo, S., Pastor, J. M., Iborra, J. L., Sauer, U., and Cánovas, M. (2012) Acetate scavenging activity in Escherichia coli: interplay of acetyl-CoA synthetase and the PEP-glyoxylate cycle in chemostat cultures. Appl. Microbiol. Biotechnol. 93, 2109 –2124 6. Lorenz, M. C., and Fink, G. R. (2001) The glyoxylate cycle is required for fungal virulence. Nature 412, 83– 86 7. Jung, J., Noh, J., and Park, W. (2011) Physiological and metabolic responses for hexadecane degradation in Acinetobacter oleivorans DR1. J. Microbiol. 49, 208 –215 8. Jung, J., and Park, W. (2013) Comparative genomic and transcriptomic analyses reveal habitat differentiation and different transcriptional responses during pectin metabolism in Alishewanella species. Appl. Environ. Microbiol. 79, 6351– 6361 9. Kim, S. J., Kweon, O., and Cerniglia, C. E. (2009) Proteomic applications to elucidate bacterial aromatic hydrocarbon metabolic pathways. Curr. Opin. Microbiol. 12, 301–309 10. Kohanski, M. A., Dwyer, D. J., Hayete, B., Lawrence, C. A., and Collins, J. J. (2007) A common mechanism of cellular death induced by bactericidal antibiotics. Cell 130, 797– 810 11. Farr, S. B., and Kogoma, T. (1991) Oxidative stress responses in Esche-

VOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response richia coli and Salmonella typhimurium. Microbiol. Rev. 55, 561–585 12. Nandakumar, M., Nathan, C., and Rhee, K. Y. (2014) Isocitrate lyase mediates broad antibiotic tolerance in Mycobacterium tuberculosis. Nat. Commun. 5, 4306 13. Rui, B., Shen, T., Zhou, H., Liu, J., Chen, J., Pan, X., Liu, H., Wu, J., Zheng, H., and Shi, Y. (2010) A systematic investigation of Escherichia coli central carbon metabolism in response to superoxide stress. BMC Syst. Biol. 4, 122 14. Vanni, P., Giachetti, E., Pinzauti, G., and McFadden, B. A. (1990) Comparative structure, function and regulation of isocitrate lyase, an important assimilatory enzyme. Comp. Biochem. Physiol. B. 95, 431– 458 15. Govan, J. R., and Deretic, V. (1996) Microbial pathogenesis in cystic fibrosis: mucoid Pseudomonas aeruginosa and Burkholderia cepacia. Microbiol. Rev. 60, 539 –574 16. Parsek, M. R., and Singh, P. K. (2003) Bacterial biofilms: an emerging link to disease pathogenesis. Annu. Rev. Microbiol. 57, 677–701 17. Boles, B. R., and Singh, P. K. (2008) Endogenous oxidative stress produces diversity and adaptability in biofilm communities. Proc. Natl. Acad. Sci. U.S.A. 105, 12503–12508 18. Díaz-Pérez, A. L., Román-Doval, C., Díaz-Pérez, C., Cervantes, C., SosaAguirre, C. R., López-Meza, J. E., and Campos-García, J. (2007) Identification of the aceA gene encoding isocitrate lyase required for the growth of Pseudomonas aeruginosa on acetate, acyclic terpenes, and leucine. FEMS Microbiol. Lett. 269, 309 –316 19. Heo, A., Jang, H. J., Sung, J. S., and Park, W. (2014) Global transcriptome and physiological responses of Acinetobacter oleivorans DR1 exposed to distinct classes of antibiotics. PLoS ONE 9, e110215 20. Lee, Y., Yeom, J., Kim, J., Jung, J., Jeon, C. O., and Park, W. (2010) Phenotypic and physiological alterations by heterologous acylhomoserine lactone synthase expression in Pseudomonas putida. Microbiology 156, 3762–3772 21. Hamada, M., Toyofuku, M., Miyano, T., and Nomura, N. (2014) cbb3-type cytochrome c oxidases, aerobic respiratory enzymes, impact the anaerobic life of Pseudomonas aeruginosa PAO1. J. Bacteriol. 196, 3881–3889 22. Kim, J., Hong, H., Heo, A., and Park, W. (2013) Indole toxicity involves the inhibition of adenosine triphosphate production and protein folding in Pseudomonas putida. FEMS Microbiol. Lett. 343, 89 –99 23. Markowitz, V. M., Chen, I. M., Chu, K., Szeto, E., Palaniappan, K., Pillay, M., Ratner, A., Huang, J., Pagani, I., Tringe, S., Huntemann, M., Billis, K., Varghese, N., Tennessen, K., Mavromatis, K., Pati, A., Ivanova, N. N., and Kyrpides, N. C. (2014) IMG/M 4 version of the integrated metagenome comparative analysis system. Nucleic Acids Res. 42, D568 —D573 24. Letunic, I., and Bork, P. (2011) Interactive Tree Of Life v2: online annotation and display of phylogenetic trees made easy. Nucleic Acids Res. 39, W475–W478 25. Kretzschmar, U., Khodaverdi, V., Jeoung, J. H., and Görisch, H. (2008) Function and transcriptional regulation of the isocitrate lyase in Pseudomonas aeruginosa. Arch. Microbiol. 190, 151–158 26. Grostern, A., Sales, C. M., Zhuang, W. Q., Erbilgin, O., and Alvarez-Cohen, L. (2012) Glyoxylate metabolism is a key feature of the metabolic degradation of 1,4-dioxane by Pseudonocardia dioxanivorans strain CB1190. Appl. Environ. Microbiol. 78, 3298 –3308 27. Akhova, A. V., and Tkachenko, A. G. (2014) ATP/ADP alteration as a sign of the oxidative stress development in Escherichia coli cells under antibiotic treatment. FEMS Microbiol. Lett. 353, 69 –76 28. Gunsalus, R. P. (1992) Control of electron flow in Escherichia coli: coordinated transcription of respiratory pathway genes. J. Bacteriol. 174, 7069 –7074 29. Singh, R., Mailloux, R. J., Puiseux-Dao, S., and Appanna, V. D. (2007) Oxidative stress evokes a metabolic adaptation that favors increased NADPH synthesis and decreased NADH production in Pseudomonas fluorescens. J. Bacteriol. 189, 6665– 6675 30. Andrae, U., Singh, J., and Ziegler-Skylakakis, K. (1985) Pyruvate and related ␣-ketoacids protect mammalian cells in culture against hydrogen peroxide-induced cytotoxicity. Toxicol. Lett. 28, 93–98 31. Wang, W., Wu, Z., Dai, Z., Yang, Y., Wang, J., and Wu, G. (2013) Glycine metabolism in animals and humans: implications for nutrition and health. Amino Acids 45, 463– 477

MAY 27, 2016 • VOLUME 291 • NUMBER 22

32. Stadtman, T. C. (1996) Selenocysteine. Annu. Rev. Biochem. 65, 83–100 33. Bryantseva, I., Gorlenko, V. M., Kompantseva, E. I., Imhoff, J. F., Süling, J., and Mityushina, L. (1999) Thiorhodospira sibirica gen. nov., sp. nov., a new alkaliphilic purple sulfur bacterium from a Siberian soda lake. Int. J. Syst. Bacteriol. 49, 697–703 34. McKinney, J. D., Höner zu Bentrup, K., Muñoz-Elías, E. J., Miczak, A., Chen, B., Chan, W. T., Swenson, D., Sacchettini, J. C., Jacobs, W. R. Jr., and Russell, D. G. (2000) Persistence of Mycobacterium tuberculosis in macrophages and mice requires the glyoxylate shunt enzyme isocitrate lyase. Nature 406, 735–738 35. Lindsey, T. L., Hagins, J. M., Sokol, P. A., and Silo-Suh, L. A. (2008) Virulence determinants from a cystic fibrosis isolate of Pseudomonas aeruginosa include isocitrate lyase. Microbiology 154, 1616 –1627 36. Wall, D. M., Duffy, P. S., Dupont, C., Prescott, J. F., and Meijer, W. G. (2005) Isocitrate lyase activity is required for virulence of the intracellular pathogen Rhodococcus equi. Infect. Immun. 73, 6736 – 6741 37. Dunn, M. F., Ramírez-Trujillo, J. A., and Hernández-Lucas, I. (2009) Major roles of isocitrate lyase and malate synthase in bacterial and fungal pathogenesis. Microbiology 155, 3166 –3175 38. Jung, J., Madsen, E. L., Jeon, C. O., and Park, W. (2011) Comparative genomic analysis of Acinetobacter oleivorans DR1 to determine strainspecific genomic regions and gentisate biodegradation. Appl. Environ. Microbiol. 77, 7418 –7424 39. Sunnarborg, A., Klumpp, D., Chung, T., and LaPorte, D. C. (1990) Regulation of the glyoxylate bypass operon: cloning and characterization of iclR. J. Bacteriol. 172, 2642–2649 40. Spiro, S., and Guest, J. R. (1991) Adaptive responses to oxygen limitation in Escherichia coli. Trends Biochem. Sci. 16, 310 –314 41. Lorca, G. L., Ezersky, A., Lunin, V. V., Walker, J. R., Altamentova, S., Evdokimova, E., Vedadi, M., Bochkarev, A., and Savchenko, A. (2007) Glyoxylate and pyruvate are antagonistic effectors of the Escherichia coli IclR transcriptional regulator. J. Biol. Chem. 282, 16476 –16491 42. Cortay, J. C., Nègre, D., Scarabel, M., Ramseier, T. M., Vartak, N. B., Reizer, J., Saier, M. H. Jr., and Cozzone, A. J. (1994) In vitro asymmetric binding of the pleiotropic regulatory protein, FruR, to the ace operator controlling glyoxylate shunt enzyme synthesis. J. Biol. Chem. 269, 14885–14891 43. Resnik, E., Pan, B., Ramani, N., Freundlich, M., and LaPorte, D. C. (1996) Integration host factor amplifies the induction of the aceBAK operon of Escherichia coli by relieving IclR repression. J. Bacteriol. 178, 2715–2717 44. Hagins, J. M., Scoffield, J. A., Suh, S. J., and Silo-Suh, L. (2010) Influence of RpoN on isocitrate lyase activity in Pseudomonas aeruginosa. Microbiology 156, 1201–1210 45. Chung, J. C., Rzhepishevska, O., Ramstedt, M., and Welch, M. (2013) Type III secretion system expression in oxygen-limited Pseudomonas aeruginosa cultures is stimulated by isocitrate lyase activity. Open Biol. 3, 120131 46. Hogardt, M., and Heesemann, J. (2013) Microevolution of Pseudomonas aeruginosa to a chronic pathogen of the cystic fibrosis lung. Curr. Top Microbiol. Immunol. 358, 91–118 47. Hentzer, M., Teitzel, G. M., Balzer, G. J., Heydorn, A., Molin, S., Givskov, M., and Parsek, M. R. (2001) Alginate overproduction affects Pseudomonas aeruginosa biofilm structure and function. J. Bacteriol. 183, 5395–5401 48. Alvarez-Ortega, C., and Harwood, C. S. (2007) Responses of Pseudomonas aeruginosa to low oxygen indicate that growth in the cystic fibrosis lung is by aerobic respiration. Mol. Microbiol. 65, 153–165 49. Chen, F., Xia, Q., and Ju, L. K. (2003) Aerobic denitrification of Pseudomonas aeruginosa monitored by online NAD(P)H fluorescence. Appl. Environ. Microbiol. 69, 6715– 6722 50. Olivares, J., Álvarez-Ortega, C., and Martinez, J. L. (2014) Metabolic compensation of fitness costs associated with overexpression of the multidrug efflux pump MexEF-OprN in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 58, 3904 –3913 51. Bedhomme, M., Zaffagnini, M., Marchand, C. H., Gao, X. H., MoslonkaLefebvre, M., Michelet, L., Decottignies, P., and Lemaire, S. D. (2009) Regulation by glutathionylation of isocitrate lyase from Chlamydomonas reinhardtii. J. Biol. Chem. 284, 36282–36291 52. Claassen, P. A. M., van den Heuvel, M. H. M. J., and Zehnder, A. J. B. (1987) Enzyme profiles of Thiobacillus versutus after aerobic and denitrifying

JOURNAL OF BIOLOGICAL CHEMISTRY

11937

Role of Glyoxylate Shunt in Oxidative Stress Response growth: regulation of isocitrate lyase. Arch. Microbiol. 147, 30 –36 53. Whittaker, J. W. (2012) Non-heme manganese catalase: the “other” catalase. Arch. Biochem. Biophys. 525, 111–120 54. Kim, E. J., Sabra, W., and Zeng, A. P. (2003) Iron deficiency leads to inhibition of oxygen transfer and enhanced formation of virulence factors in cultures of Pseudomonas aeruginosa PAO1. Microbiology 149, 2627–2634 55. Wiens, J. R., Vasil, A. I., Schurr, M. J., and Vasil, M. L. (2014) Iron-regulated expression of alginate production, mucoid phenotype, and biofilm

11938 JOURNAL OF BIOLOGICAL CHEMISTRY

formation by Pseudomonas aeruginosa. mBio 5, e01010 – e01013 56. Rock, J. D., and Moir, J. W. (2005) Microaerobic denitrification in Neisseria meningitidis. Biochem. Soc. Trans. 33, 134 –136 57. Lawley, P. D., and Thatcher, C. J. (1970) Methylation of deoxyribonucleic acid in cultured mammalian cells by N-methyl-N⬘-nitro-N-nitrosoguanidine. The influence of cellular thiol concentrations on the extent of methylation and the 6-oxygen atom of guanine as a site of methylation. Biochem. J. 116, 693–707

VOLUME 291 • NUMBER 22 • MAY 27, 2016

Role of Glyoxylate Shunt in Oxidative Stress Response.

The glyoxylate shunt (GS) is a two-step metabolic pathway (isocitrate lyase, aceA; and malate synthase, glcB) that serves as an alternative to the tri...
2MB Sizes 1 Downloads 6 Views