Abstract
The Arc two-component system, comprising the ArcB sensor kinase and the ArcA response regulator, modulates the expression of numerous genes in response to respiratory growth conditions. Under aerobic growth conditions, the ubiquinone electron carriers were proposed to silence the kinase activity of ArcB by oxidizing two cytosol-located redox-active cysteine residues that participate in intermolecular disulfide bond formation. Here, we confirm the role of the ubiquinone electron carriers as the silencing signal of ArcB in vivo, we show that the redox potential of ArcB is about −41 mV, and we demonstrate that the menaquinols are required for proper ArcB activation upon a shift from aerobic to anaerobic growth conditions. Thus, an essential link in the Arc signal transduction pathway connecting the redox state of the quinone pool to the transcriptional apparatus is elucidated.
INTRODUCTION
The Arc (anoxic redox control) two-component system (TCS) is a key element in the transcriptional regulatory network that allows facultative anaerobic bacteria to sense various respiratory growth conditions and adapt their gene expression accordingly (1–4). This system comprises the cytoplasmic response regulator ArcA and the membrane-anchored sensor kinase ArcB (5, 6). ArcA is a typical response regulator possessing an N-terminal receiver domain with a conserved Asp residue at position 54 and a C-terminal helix-turn-helix DNA binding domain. In contrast, ArcB is an unorthodox sensor kinase, having a very short periplasmic sequence of only 16 amino acid residues delimited by two canonical transmembrane segments. Interestingly, these segments of ArcB do not directly participate in signal sensing but, rather, serve as an anchor that keeps the protein close to the source of the signal (7). Moreover, ArcB contains three catalytic domains: an N-terminal transmitter domain with a conserved His292 residue, a central receiver domain with a conserved Asp576 residue, and a C-terminal phosphotransfer domain with a conserved His717 residue (6, 8). Finally, in the linker, that is, the region connecting the catalytic domains with the transmembrane domain, there are a functional leucine zipper (9) and a PAS domain (10).
Under reducing conditions, ArcB autophosphorylates through an intramolecular reaction (11), a process shown to be enhanced by certain anaerobic metabolites, such as d-lactate, acetate, and pyruvate (12, 13), and transphosphorylates ArcA via a His292 → Asp576 → His717 → Asp54 phosphorelay (14, 15). Phosphorylated ArcA (ArcA-P), in turn, represses the expression of many operons involved in respiratory metabolism and activates others encoding proteins involved in fermentative metabolism (16–19). Under nonstimulating conditions, ArcB acts as a specific ArcA-P phosphatase that catalyzes the dephosphorylation of ArcA-P by a reverse Asp54 → His717 → Asp576 → Pi phosphorelay (20, 21). Previously, it was reported that regulation of the catalytic activity of ArcB is set by rotational movements that alter the orientation of the cytosolic portion of ArcB (22) and that the molecular event for ArcB regulation involves the oxidation of two cytosol-located redox-active cysteine residues that participate in intermolecular disulfide bond formation, a reaction in which ubiquinones (UQ) act as direct oxidants (23, 24). However, in a recent study, it was reported that the aerobic and anaerobic levels of an ArcA-P-specific reporter (cydA−176+1-lacZ) in a mutant blocked in ubiquinone biosynthesis were similar to the ones of the isogenic wild-type (wt) strain, leading the authors to conclude that aerobic inhibition of the ArcB kinase in vivo can be due to other effectors and thus is regulated not only by ubiquinone (25). On the other hand, a mutant strain blocked in the menaquinone (MQ) biosynthetic pathway resulted in the loss of anaerobic activation of reporter expression, suggesting that the menaquinone pool plays a major role in ArcB redox regulation (25). Moreover, in another recent study, it was reported that no correlation between ArcA activity/phosphorylation and the ubiquinone/ubiquinol redox profile was found (26). It was therefore suggested that ArcB integrates the response to multiple signals (e.g., ubiquinone, menaquinone, acetate, and other fermentation products) and that the rate of fermentation product synthesis exerts a greater influence on ArcA activity than the redox state of the quinone pools (26).
Here we present experiments addressing the above-mentioned questions and discrepancies. Our results confirm our previous findings that ubiquinones (E′° = +100 mV) are necessary for silencing the kinase activity of ArcB during aerobic growth or upon a shift from anaerobiosis to aerobiosis and also extend these studies by determining the redox potential of ArcB to be about −41 mV and demonstrating that the menaquinone electron carriers (E′° = −74 mV) are required for activation of ArcB upon a shift to anoxic conditions. A simple model for the ArcB redox signaling which differs significantly from those previously suggested (25, 26) is presented and discussed.
MATERIALS AND METHODS
Strain and plasmid constructions.
The strains and plasmids used in this work are listed in Table 1. Strain IFC5006 (ΔmenFDHB::Kanr Δfnr::Tn9-Chlr cydA′-lacZ) was constructed by homologous recombination using the lambda red recombinase system (27). Briefly, a PCR-amplified fragment, using primers menF-up-pKD-Fw and menB-down-pKD-Rv (Table 2) and plasmid pKD4 (27) as the template, was used to replace the menFDHB operon with a kanamycin (Kan) cassette in strain ECL5003 (Δfnr::Tn9-Chlr cydA′-lacZ) (7). Strain IFC5007 (sdh′-lacZ) was constructed by P1vir transduction of the Φ(sdh′-lacZ) allele from strain ECL594 (28) into strain MC4100 (29). Strain IFC5008 (ΔubiCA::Kanr sdh′-lacZ) was constructed by P1vir transduction of the ΔubiCA::Kanr allele from strain ECL5039 (ΔubiCA::Kanr cydA′-lacZ) (23) into strain IFC5007. Strain IFC5009 (ΔubiC::Kanr) was constructed by P1vir transduction of the ΔubiC::Kanr allele from strain JW5713 (30) into strain ECL5001. The relevant genotype of all mutants was verified by PCR. To construct plasmid pMX538, the complete men operon (menFDHBCE), including its promoter region, was PCR amplified, using primers menF-Fw and menE-Rv (Table 2) and chromosomal DNA from MC4100 as the template, and cloned into the BamHI and HindIII sites of plasmid pBR322 (31). To construct plasmid pMX537, the ubiCA operon and its promoter region were PCR amplified, using primers ubiC-Fw and ubiA-Rv (Table 2) and chromosomal DNA from MC4100 as the template, and cloned into the EcoRI and HindIII sites of the low-copy-number plasmid pEXT21 (32). Construction of plasmids pQE30ArcB78-778 and pMX517 has been described earlier (9, 15).
Table 1.
Strain or plasmid | Relevant characteristicsa | Source |
---|---|---|
Strains | ||
MC4100 | F− araD139 (argF-lac) U169 rpsL150 relA1 flbB5301 deoC ptsF25 rbsR | 29 |
BW25113 | K-12 wild type | 30 |
JW5713 | BW25113 ΔubiC::Kanr | 30 |
ECL5001 | MC4100 λΦ(cydA′-lacZ) | 7 |
ECL5039 | ECL5001 ΔubiCA::Kanr | 23 |
ECL5003 | MC4100 Δfnr::Tn9(Chlr) λΦ(cydA′-lacZ) | 7 |
IFC5006 | ECL5003 ΔmenFDHB::Kanr | This work |
ECL594 | MC4100 sdh+ Φ(sdh-lacZ) arcBl zgi::Tn10 | 28 |
IFC5007 | MC4100 Φ(sdh-lacZ) from ECL594 | This work |
IFC5008 | IFC5007 ΔubiCA::Kanr from ECL5039 | This work |
IFC5009 | ECL5001 ΔubiC::Kanr from JW5713 | This work |
Plasmids | ||
pEXT21 | Low-copy-no. vector; Sptr | 32 |
pBR322 | Cloning vector; Ampr Tetr | 31 |
pMX517 | Full-length arcB under the control of the ara promoter; Ampr | 9 |
pQE30ArcB78-778 | ArcB78-778 in pQE30 under an IPTG-inducible promoter | 15 |
pMX537 | pEXT21::ubiCA; Sptr | This work |
pMX538 | pBR322::menFDHBCE; Ampr | This work |
Kanr, kanamycin resistant; Chl, chloramphenicol; Spt, spectinomycin; Amp, ampicillin; Tet, tetracycline.
Table 2.
Name | Sequence (5′–3′)a | Use |
---|---|---|
menF-up-pKD-Fw | CCCCGTATAATGTGAGGCTTTTAACAGGGAGAGGTCCGCGTGTAGGCTGGAGCTGCTTC | Forward primer for men operon deletion |
menB-down-pKD-Rv | CGTCCATGGGGATCTGCCAGCGGTATACCTGCGCGCTACGCATATGAATATCCTCCTTAG | Reverse primer for men operon deletion |
menF-Fw | CCCGGATCCGTGATACACGTATCGATG (BamHI) | Forward primer for men operon cloning and men mutant confirmation |
menE-Rv | CCCAAGCTTGACCAGCCATTCCATTGC (HindIII) | Reverse primer for men operon cloning and men mutant confirmation |
ubiC-Fw | GATACCCAACAGATGATCG | Forward primer for ubi operon cloning and ubi mutant confirmation |
ubiA-Rv | CCCAAGCTTCGCACGCTACAGCTGC (HindIII) | Reverse primer for ubi operon cloning and ubi mutant confirmation |
k1-pKD4 | GTCATAGCCGAATAGCC | Primer for mutant genotype confirmation |
k2-pKD4 | GTGCCCTGAATGAACTG | Primer for mutant genotype confirmation |
Restriction enzymes whose sites were introduced for subsequent cloning (underlined) are indicated in parentheses.
Growth conditions.
Escherichia coli strains were routinely grown in Luria-Bertani (LB) medium at 37°C. For β-galactosidase activity assays, either LB broth buffered with 0.1 M MOPS (morpholinepropanesulfonic acid; pH 7.4) and supplemented with 20 mM d-xylose or defined minimal medium [1 mM KH2PO4, 40 mM KCl, 34 mM NaCl, 20 mM (NH4)2SO4, 1 μM FeSO4, 0.3 mM MgSO4, 1 μM ZnCl2, 10 μM CaCl2, and 0.1 M MOPS; pH 7.4] supplemented with 0.4% (wt/vol) glucose and 15 mg/liter of thiamine was used. When necessary, ampicillin, kanamycin, tetracycline, or spectinomycin was used at a final concentration of 100, 50, 20, or 50 μg ml−1, respectively. l-Arabinose and isopropyl-β-d-thiogalactopyranoside (IPTG) were used at final concentrations of 0.13 and 1 mM to induce expression of full-length ArcB and ArcB78-778, respectively. For complementation assays, 1,4-dihydroxy-2-naphtoic acid (DHNA) was added to the medium at a final concentration of 5 μM.
ArcB-enriched inverted vesicle preparation, purification of His6-tagged proteins, and phosphorylation assays.
ArcB-enriched inverted membrane vesicles and His6-tagged ArcB78-778, used in phosphorylation assays, were prepared as described previously (9, 15, 33). Phosphorylation assays were carried out at room temperature in the presence of 40 mM [γ-32P]ATP (specific activity, 2 Ci mmol−1; New England Nuclear), 33 mM HEPES (pH 7.5), 50 mM KCl, 5 mM MgCl2, 0.1 mM EDTA, and 10% glycerol. For the estimation of the ArcB redox potential, quinone analogues, including 1,4-benzoquinone, 1,2-naphtoquinone, 1,4-naphtoquinone, juglone, menadione, plumbagin, lawsone, and anthraquinone-2-sulfonate, were used at 1 mM. The phosphorylation reactions were initiated by the addition of [γ-32P]ATP and terminated 2.5 min later by addition of an equal volume of 4× SDS sample buffer, and the mixtures were immediately subjected to SDS-PAGE on 10% polyacrylamide gels. The radioactivity of proteins resolved in the gels was analyzed by using a PhosphorImager (Molecular Dynamics).
Determination of the redox potential of the disulfide bonds of ArcB.
The redox potential value of ArcB was determined using cysteine/cystine couple redox buffers as previously described (34) but with some modifications. Briefly, ArcB-enriched inverted membrane vesicles containing ∼1 μM concentrations of the full-length ArcB were incubated for 3 h at 25°C in the redox buffers (33 mM HEPES [pH 7.5], 50 mM KCl, 5 mM MgCl2, 0.1 mM EDTA, 10 μM UQ0, 40 μM cystine, and different concentrations of l-cysteine, ranging from 0.04 μM to 1 mM). All solutions were flushed with nitrogen to eliminate oxygen and minimize air oxidation. After incubation, samples were subjected to phosphorylation by addition of 40 mM [γ-32P]ATP and analyzed as described above. The results were expressed as the fractions of reduced (active) protein, and the equilibrium constant (Keq) was calculated according to the equation R = ([Cys]2/[cystine])/{Keq + ([Cys]2/[cystine])}, where R is a fraction of the reduced proteins at equilibrium. The standard redox potential of ArcB was calculated from the Nernst equation, E′°ArcB = E′°Cys/Cystine − (RT/2F) ln Keq, where F and R are the Faraday constant and the gas constant, respectively, using a value of −200 mV for the Cys/cystine redox couple (E′°Cys/Cystine) (35, 36).
β-Galactosidase activity assay.
For the aerobiosis-to-anaerobiosis shift, cells were cultured in 50 ml of medium in 250-ml baffled flasks at 37°C with shaking (300 rpm). At an optical density at 600 nm (OD600) of 0.2, part of the culture was transferred to five prewarmed screw-cap tubes, filled up to the rim, and stirred by a magnet. The rest of the aerobic culture was further incubated with shaking at the same temperature. The time course of the experiment was followed by taking a sample from a filled screw-cap tube (anaerobiosis) and from the baffled flasks (aerobiosis) at each chosen time. For the anaerobiosis-to-aerobiosis shift, cells were cultured in seven screw-cap tubes filled with medium up to the rim at 37°C and stirred by a magnet. At an OD600 of 0.2, the content of a tube was passed to a 250-ml prewarmed baffled flask and incubated with shaking at the same temperature. Samples from aerobic and anaerobic cultures (the rest of the screw-cap tubes) were withdrawn at 0, 20, 40, 60, 90, and 120 min after the shift. β-Galactosidase activity was assayed and expressed in Miller units as described previously (37).
RESULTS
The ubiquinone electron carriers are required for ArcB silencing upon a shift from anaerobic to aerobic growth conditions.
It has been previously reported that under aerobic conditions, the ubiquinone electron carriers act as negative signals that silence the kinase activity of ArcB (23) by oxidizing two cytosol-located redox-active cysteine residues that participate in intermolecular disulfide bond formation (24). However, in a recent study, the role of ubiquinones as the specific silencing signal for ArcB was disputed, as it was found that the aerobic and anaerobic levels of an ArcA-P-specific reporter in a ubiC mutant (blocked in the ubiquinone biosynthesis) were similar to the ones of the isogenic wild-type strain (25). To address this discrepancy of whether or not ubiquinones are required for ArcB regulation in vivo, strain ECL5039, a ubiCA mutant carrying the ArcA-P activatable cydA′-lacZ operon fusion (23), and its isogenic wild type were grown anaerobically in buffered Luria broth, and at an OD600 of ∼0.2, the culture was divided in two. One part was shifted to aerobiosis, whereas the other half was kept under anaerobiosis as a control, and the time course of the β-galactosidase activity was followed (Fig. 1). Monitoring the Arc activity during a shift from stimulatory to nonstimulatory growth conditions in buffered LB was decided because ArcB has been shown to respond rapidly to the presence or absence of the specific stimulus (21) and because the ubiCA mutant grows poorly aerobically (38–40). It was found that shifting the anaerobic culture to aerobiosis caused a rapid decrease in reporter expression in the wild-type strain, indicative of ArcB/A silencing (Fig. 1A). However, no decrease in reporter expression was observed in the ubiCA mutant during the course of the experiment (Fig. 1B), indicating a defect in the inhibition of the ArcB kinase activity and thereby no silencing of the system. On the other hand, complementation of the mutant strain with a low-copy-number plasmid carrying the ubiCA operon (pMX537) fully restored reporter expression to wild-type levels (Fig. 1C). To ensure that the above-described result is due to the lack of inhibition of the ArcB kinase activity and not specific to the cydA′-lacZ reporter, we pursued a similar experiment using strains IFC5007 (wt) and IFC5008 (ubiCA), carrying the ArcA-P repressible sdh-lacZ operon fusion. The two isogenic strains were grown anaerobically to an OD600 of ∼0.2, and half of the cultures were shifted to aerobiosis for 2 h. It was found that shifting the anaerobic cultures to aerobiosis resulted in release of the ArcB-dependent repression of sdh-lacZ expression in the wild-type strain but not in the ubiCA mutant (Fig. 1D), confirming the above-mentioned result. Therefore, it can be concluded that the ubiquinone electron carriers are required for rapid silencing of the ArcB kinase activity upon a shift from anoxic to oxic growth conditions.
We next attempted to discern whether the differences between the above-mentioned results and the ones reported by Bekker et al. (25) are due to the distinct experimental settings. Strain ECL5039 (ubiCA) and its isogenic wild-type strain ECL5001, both carrying the cydA′-lacZ reporter, were grown aerobically and anaerobically in defined minimal medium with glucose as the sole carbon source, and at mid-exponential growth, the β-galactosidase activity was measured (Fig. 1E). As expected, the anaerobic reporter expression was 5-fold higher than the aerobic expression in the wild-type strain. On the other hand, the aerobic reporter expression in the ubiCA mutant strain was 2.7-fold higher than that of the wt, indicative of diminished inhibition of ArcB kinase activity, in agreement with the above-presented results and our previous report (23). Moreover, the anaerobic reporter expression in the ubiCA mutant strain was only 1.3-fold higher than the aerobic expression, revealing a significant loss of regulation due to the elevated ArcB kinase activity under aerobic conditions. Thus, further support for the role of ubiquinones in the inhibition of the ArcB kinase activity, in agreement with our previous observations (23), is provided.
Finally, we tested whether the above-mentioned discrepancy is due to the different strains used in the two studies. It has been previously demonstrated that ubiquinone-deficient mutants grow poorly aerobically and that they fail to grow on succinate-mineral salt medium (38). Therefore, we compared the aerobic growth of ECL5039 (the ubiCA mutant used in this study), JW5713 (the ubiC mutant strain used by Bekker et al.), and their isogenic wild-type strains (ECL5001 and BW25113, respectively) on glucose- and succinate-mineral salt medium-agar plates. Curiously, although ECL5039 grew poorly on glucose and failed to grow on succinate, strain JW5713 grew almost equally as well as the wild-type strain on both substrates (see Fig. S1 in the supplemental material), indicating that a possible mutation suppresses the ubi mutant growth phenotype. Moreover, we found that this trait of JW5713 was cotransducible with the ubi::kan mutation by P1vir (see Fig. S1 in the supplemental material). Therefore, it is possible that the different strains used in the two studies might contribute to differences in ArcB activity and subsequent control of reporter expression. However, no further efforts were made to map this mutation because this was outside the scope of the present study.
The redox potential of ArcB is −41 mV.
The fact that the molecular event for ArcB regulation involves the oxidation/reduction of cysteine residues C180 and C241 renders the assessment of the ArcB redox potential of foremost importance. Therefore, in an effort to approximate the ArcB redox potential, the effect of several quinone analogs, covering a broad range of redox potential values, on the [γ32P] ATP-dependent phosphorylation of ArcB was tested (Fig. 2A). The analogues tested were 1,4-benzoquinone (E′° = +274 mV), 1,2-naphtoquinone (E′° = +134 mV), 1,4-naphtoquinone (E′° = +69 mV), juglone (E′° = +30 mV), menadione (E′° = ±0 mV), plumbagin (E′° = −29 mV), lawsone (E′° = −137 mV), and anthraquinone-2-sulfonate (E′° = −225mV). It was found that quinone analogs having a positive redox potential were effective inhibitors of the kinase activity of ArcB whereas analogs with negative redox potentials were without effect. The same effect was obtained with the quinone analogs for both purified ArcB78-778 and ArcB-enriched inverted membrane vesicles (Fig. 2A).
To determine a more accurate redox potential for ArcB, we took advantage of the fact that the exact redox potential of a cysteine pair in a protein can be achieved by quantifying the steady-state ratios of their reduced and oxidized forms in the presence of excess of a composition-defined redox buffer. Therefore, we determined the redox potential of full-length ArcB embedded in inverted membrane vesicles from its redox equilibrium constant (Keq) with a cysteine/cystine redox buffer at 25°C and using the ArcB phosphorylation ability as the indicator for the reduced/oxidized state of ArcB (Fig. 2B). The percentage of net phosphorylation as a function of the [cysteine]2/[cystine] relation was used for the calculations of the Keq value and the standard redox potential of ArcB, which were found to be 4.2 × 10−6 M and −41 mV, respectively (Fig. 2B).
Thus, the ubiquinones, having a redox potential of +100 mV, may serve perfectly as the ArcB oxidizing agent, in accordance with the above-presented in vivo results.
The menaquinone electron carriers are required for reactivation of ArcB upon a shift from aerobic to anaerobic growth conditions.
Given that the redox potential of ArcB was found to be −41 mV, significantly higher than the redox potential of menaquinone (−74 mV), we considered that menaquinols may serve as the specific electron donors for ArcB. In such a scenario, the electrons would flow from menaquinols to ArcB, resulting in the reduction of the cysteine residues and activation of the kinase activity of ArcB. We therefore argued that if menaquinones were indeed involved in the regulation of ArcB, no activation of cydA′-lacZ expression should be observed in a menaquinone-deficient mutant after a shift from aerobic to anaerobic growth conditions. To test this, a mutant strain blocked in the menaquinone biosynthetic pathway was constructed (IFC5006) by deleting part of the men operon (menF, menD, menH, and menB) in a cydA′-lacZ-carrying strain. The choice of blocking the first steps of this pathway was in order to prevent the undesirable accumulation of intermediates. The mutant and its isogenic wild-type strain were grown aerobically in buffered Luria broth, and at an OD600 of ∼0.2, the culture was divided in two. One part was kept under aerobic conditions as a control, whereas the other was shifted to anaerobiosis, and the time course of the β-galactosidase activity was followed (Fig. 3). It was found that shifting the aerobic culture to anaerobiosis caused a rapid increase in reporter expression in the wild-type strain (Fig. 3A), indicative of proper activation of the ArcB kinase activity. In contrast, no increase in reporter expression was observed in the menFDHB mutant (Fig. 3B), indicating a failure in the activation of ArcB kinase activity. Reporter expression was restored to wild-type levels when this mutant strain was complemented with a low-copy-number plasmid carrying the menFDHBCE operon (pMX538) (Fig. 3C). Likewise, addition of 1,4-dihydroxy-2-naphtoic acid (DHNA), which is the product of MenB and substrate of MenA to produce demethylmenaquinone, which is converted to menaquinone by UbiE, to the growth medium of the menFDHB mutant did not affect the aerobic expression of cydA′-lacZ but fully restored reporter expression after a shift to anaerobiosis (Fig. 3D), in agreement with a previous report (25). Therefore, it can be concluded that the menaquinone electron carriers are required for proper reactivation of the ArcB kinase activity after a shift from aerobic to anaerobic growth conditions.
DISCUSSION
The Arc two-component system is a complex signal transduction system that plays a key role in regulating energy metabolism at the level of transcription in bacteria. Under reducing growth conditions, ArcB autophosphorylates and transphosphorylates ArcA, which, in turn, represses or activates the expression of its target operons. Under aerobic conditions, ArcB is inactivated as a kinase but acts as a phosphatase that catalyzes the dephosphorylation of ArcA-P, thereby releasing its transcriptional regulation. The molecular mechanism of ArcB kinase silencing was shown to require the ubiquinone-dependent oxidation of two cysteine residues, C180 and C241, leading to intermolecular disulfide bond formation (23, 24). In this study, we demonstrate that ubiquinones (E′° = +100 mV) are required for ArcB silencing, in accordance with our previous report (23). In addition, we determine the midpoint redox potential of ArcB to be approximately −41 mV, and we demonstrate, in agreement with a previous study (25), that menaquinols (E′° = −74 mV) are required for proper ArcB reactivation upon a shift from aerobic to anaerobic growth conditions. Our results led us to propose a revised model for Arc redox signaling (Fig. 4) which differs significantly from those previously suggested (25, 26) but is in agreement with the model proposed by Portnoy and coworkers (41). Our results are also in accordance with the previous findings that despite the aerobic growth conditions, an ArcA-P expression profile was observed in mutant strains lacking both cytochrome bo and cytochrome bd terminal oxidases (42), the ubiG gene (43), or the three terminal cytochrome oxidase genes (cydAB, cyoABCD, and cbdAB) and a quinol monooxygenase gene (ygiN) (41). These observations are promptly explained by the demonstration that these mutations result in a shift in the content of the quinone pool from the ubiquinones to menaquinones (41). Such a shift should lead to the activation of ArcB, the rise of ArcA-P levels, and the adjustment of the expression of the target operons accordingly.
Nevertheless, two recent studies (25, 26) put into question the importance of the ubiquinone pool in the regulation of the ArcB sensor kinase and our previously proposed model of ArcB regulation (3, 23, 24). In both studies, in which chemostat cultures were used, the challenge of the importance of ubiquinones in the regulation of ArcB kinase activity was based on the failure of establishing a correlation between reporter expression or the level of ArcA-P and the redox state of the ubiquinone pool (25, 26). Also, results showed that the aerobic and anaerobic expression levels of an ArcA-P-specific reporter (cyd−176+1-lacZ) in a ubiquinone-deficient mutant strain were similar to the ones of the isogenic wild-type strain (25). Of note, the ArcA activity profiles and the quinol/quinone ratios differed significantly in these two studies, although both used steady-state chemostat cultures. Nevertheless, the correlation between reporter expression and ubiquinone/menaquinone levels prompted Bekker and coauthors to propose the following model (25). Upon a shift from anaerobiosis to low aerobiosis, the menaquinone pool oxidizes rapidly, resulting in oxidation of the regulatory cysteine residues and inactivation of ArcB. A further increase in aerobiosis leads to an increase in ubiquinol, resulting in the reduction of the cysteines and, thereby, reactivation of ArcB. Finally, in complete aerobiosis, the content of the quinone pool decreases, resulting in the oxidation of the cysteines and inactivation of ArcB. However, such a model implies that both menaquinones and ubiquinones are able to silence the ArcB kinase activity whereas both menaquinols and ubiquinols are able to activate it. This can be possible only if quinone/quinol binding was sufficient for regulation of the ArcB activity but not if electron transfer between the quinones/quinols and the two redox-active cysteines of ArcB were required. An alternative explanation might be that in JW5713, a ubiC mutant strain from the Keio collection (30), the amount of demethylmenaquinone (E′° = +36 mV) increases. In such a case, demethylmenaquinone may, at least in part, compensate the absence of ubiquinone because its redox potential is considerably higher than that of ArcB (E′° = −41 mV).
On the other hand, Rofle and coworkers, using transcriptomic analysis and direct measurement of ArcA-P across the aerobiosis range (0 to 100%), reported a linear decrease in the ArcB/A activity with increasing aerobiosis (26), in agreement with our results presented here and with previous reported studies (4, 5, 16). However, no correlation between ArcB/A activity and the redox state of the ubiquinone pool was found. Therefore, the authors concluded that inhibition of ArcB phosphatase activity mediated by fermentation products, such as acetate, may be the dominant mechanism for regulating ArcB and, thereby, ArcA activity instead of the redox state of the ubiquinone pool (26). However, the 2-fold difference in total ubiquinone between anaerobiosis and full aerobiosis contrasts the ∼5-fold increase reported earlier (25, 39, 41, 44, 45) and weakens the above-mentioned conclusion. Moreover, this mode of ArcB regulation has been discarded in earlier in vivo and in vitro studies that demonstrate that fermentation products, such as acetate and lactate, have no effect on the ArcB phosphatase activity (12) and do not act as direct signals for activation of ArcB but rather act as physiologically significant effectors that amplify the ArcB kinase activity (13).
In view of the data reported herein, a revised model is presented (Fig. 4). E. coli expresses branched electron transfer chains that can deal with various electron donors and several alternative electron acceptors. A central component in these networks is the quinone pool that acts as a redox mediator. During aerobic growth, the major quinone is ubiquinone (E′° = +100 mV), which is replaced by menaquinone (E′° = −74 mV) in anaerobic growth (25, 39, 41, 44, 45). Thus, during aerobic growth conditions or upon a shift from anaerobic to aerobic growth, the predominant ubiquinone pool acts to oxidize the cysteine residues of ArcB (E′° ≈ −41 mV), resulting in disulfide bond formation and immediate silencing of the ArcB kinase activity, leading to inactivation of ArcA. On the other hand, during microaerobic growth conditions or upon a shift from aerobic to anaerobic growth, the ubiquinone pool is gradually replaced by menaquinones that act to reduce the cysteine residues of ArcB, resulting in disulfide bond breakage, activation of the ArcB kinase activity, and consequent activation of ArcA. Based on the proposed model, we predict that both UQ and MQ are competing for a single quinone-binding pocket in ArcB that is modulated by the UQ/MQ ratio under aerobic, microaerobic, and anaerobic growth conditions. Such a mechanism allows rapid responses of bacterial cells to their environment, as the respiratory growth conditions are instantly sensed and transduced into the transcriptional apparatus.
Supplementary Material
ACKNOWLEDGMENTS
We thank Georges Dreyfus and Diego Gonzalez Halphen for stimulating discussions and for critically reading the manuscript, the Unidad de Biología Molecular from the Instituto de Fisiología Celular, Universidad Nacional Autónoma de México, for oligonucleotide synthesis and sequencing, and the NBRP-E. coli at NIG for strains BW25113 and JW5713.
This work was partially supported by grant 178033 from the Consejo Nacional de Ciencia y Tecnología (CONACyT) and grant IN206412 from DGAPA, UNAM. A.F.A. thanks DGAPA-UNAM for the award of a postdoctoral fellowship.
Footnotes
Published ahead of print 3 May 2013
Supplemental material for this article may be found at http://dx.doi.org/10.1128/JB.00406-13.
REFERENCES
- 1. Georgellis D, Kwon O, Lin EC, Wong SM, Akerley BJ. 2001. Redox signal transduction by the ArcB sensor kinase of Haemophilus influenzae lacking the PAS domain. J. Bacteriol. 183:7206–7212 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Jung WS, Jung YR, Oh DB, Kang HA, Lee SY, Chavez-Canales M, Georgellis D, Kwon O. 2008. Characterization of the Arc two-component signal transduction system of the capnophilic rumen bacterium Mannheimia succiniciproducens. FEMS Microbiol. Lett. 284:109–119 [DOI] [PubMed] [Google Scholar]
- 3. Malpica R, Sandoval GR, Rodriguez C, Franco B, Georgellis D. 2006. Signaling by the Arc two-component system provides a link between the redox state of the quinone pool and gene expression. Antioxid. Redox Signal. 8:781–795 [DOI] [PubMed] [Google Scholar]
- 4. Lynch AS, Lin EC. 1996. Regulation of gene expression in Escherichia coli, p 1526–1538 In Neidhardt FC, Curtis R, Ingraham AL, Lin ECC, Low KB, Magasanik B, Reznikoff WS, Riley M, Schaechter M, Umbarger HE. (ed), Escherichia coli and Salmonella: cellular and molecular biology. American Society of Microbiology, Washington, DC [Google Scholar]
- 5. Iuchi S, Lin EC. 1988. arcA (dye), a global regulatory gene in Escherichia coli mediating repression of enzymes in aerobic pathways. Proc. Natl. Acad. Sci. U. S. A. 85:1888–1892 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Iuchi S, Matsuda Z, Fujiwara T, Lin EC. 1990. The arcB gene of Escherichia coli encodes a sensor-regulator protein for anaerobic repression of the arc modulon. Mol. Microbiol. 4:715–727 [DOI] [PubMed] [Google Scholar]
- 7. Kwon O, Georgellis D, Lynch AS, Boyd D, Lin EC. 2000. The ArcB sensor kinase of Escherichia coli: genetic exploration of the transmembrane region. J. Bacteriol. 182:2960–2966 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Ishige K, Nagasawa S, Tokishita S, Mizuno T. 1994. A novel device of bacterial signal transducers. EMBO J. 13:5195–5202 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Nuñez Oreza LA, Alvarez AF, Arias-Olguín II, Torres Larios A, Georgellis D. 2012. The ArcB leucine zipper domain is required for proper ArcB signaling. PLoS One 7:e38187. 10.1371/journal.pone.0038187 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Zhulin IB, Taylor BL, Dixon R. 1997. PAS domain S-boxes in Archaea, Bacteria and sensors for oxygen and redox. Trends Biochem. Sci. 22:331–333 [DOI] [PubMed] [Google Scholar]
- 11. Peña-Sandoval GR, Georgellis D. 2010. The ArcB sensor kinase of Escherichia coli autophosphorylates by an intramolecular reaction. J. Bacteriol. 192:1735–1739 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Georgellis D, Kwon O, Lin EC. 1999. Amplification of signaling activity of the Arc two-component system of Escherichia coli by anaerobic metabolites. An in vitro study with different protein modules. J. Biol. Chem. 274:35950–35954 [DOI] [PubMed] [Google Scholar]
- 13. Rodriguez C, Kwon O, Georgellis D. 2004. Effect of d-lactate on the physiological activity of the ArcB sensor kinase in Escherichia coli. J. Bacteriol. 186:2085–2090 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Kwon O, Georgellis D, Lin EC. 2000. Phosphorelay as the sole physiological route of signal transmission by the arc two-component system of Escherichia coli. J. Bacteriol. 182:3858–3862 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Georgellis D, Lynch AS, Lin EC. 1997. In vitro phosphorylation study of the Arc two-component signal transduction system of Escherichia coli. J. Bacteriol. 179:5429–5435 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Gunsalus RP, Park SJ. 1994. Aerobic-anaerobic gene regulation in Escherichia coli: control by the ArcAB and Fnr regulons. Res. Microbiol. 145:437–450 [DOI] [PubMed] [Google Scholar]
- 17. Liu X, De Wulf P. 2004. Probing the ArcA-P modulon of Escherichia coli by whole genome transcriptional analysis and sequence recognition profiling. J. Biol. Chem. 279:12588–12597 [DOI] [PubMed] [Google Scholar]
- 18. Lynch AS, Lin EC. 1996. Transcriptional control mediated by the ArcA two-component response regulator protein of Escherichia coli: characterization of DNA binding at target promoters. J. Bacteriol. 178:6238–6249 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Alvarez AF, Malpica R, Contreras M, Escamilla E, Georgellis D. 2010. Cytochrome d but not cytochrome o rescues the toluidine blue growth sensitivity of arc mutants of Escherichia coli. J. Bacteriol. 192:391–399 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Georgellis D, Kwon O, De Wulf P, Lin EC. 1998. Signal decay through a reverse phosphorelay in the Arc two-component signal transduction system. J. Biol. Chem. 273:32864–32869 [DOI] [PubMed] [Google Scholar]
- 21. Peña-Sandoval GR, Kwon O, Georgellis D. 2005. Requirement of the receiver and phosphotransfer domains of ArcB for efficient dephosphorylation of phosphorylated ArcA in vivo. J. Bacteriol. 187:3267–3272 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Kwon O, Georgellis D, Lin EC. 2003. Rotational on-off switching of a hybrid membrane sensor kinase Tar-ArcB in Escherichia coli. J. Biol. Chem. 278:13192–13195 [DOI] [PubMed] [Google Scholar]
- 23. Georgellis D, Kwon O, Lin EC. 2001. Quinones as the redox signal for the Arc two-component system of bacteria. Science 292:2314–2316 [DOI] [PubMed] [Google Scholar]
- 24. Malpica R, Franco B, Rodriguez C, Kwon O, Georgellis D. 2004. Identification of a quinone-sensitive redox switch in the ArcB sensor kinase. Proc. Natl. Acad. Sci. U. S. A. 101:13318–13323 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Bekker M, Alexeeva S, Laan W, Sawers G, de Mattos JT, Hellingwerf K. 2010. The ArcBA two-component system of Escherichia coli is regulated by the redox state of both the ubiquinone and the menaquinone pool. J. Bacteriol. 192:746–754 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Rolfe MD, Ter Beek A, Graham AI, Trotter EW, Asif HM, Sanguinetti G, de Mattos JT, Poole RK, Green J. 2011. Transcript profiling and inference of Escherichia coli K-12 ArcA activity across the range of physiologically relevant oxygen concentrations. J. Biol. Chem. 286:10147–10154 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Datsenko KA, Wanner BL. 2000. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc. Natl. Acad. Sci. U. S. A. 97:6640–6645 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Iuchi S, Cameron DC, Lin EC. 1989. A second global regulator gene (arcB) mediating repression of enzymes in aerobic pathways of Escherichia coli. J. Bacteriol. 171:868–873 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Casadaban MJ. 1976. Transposition and fusion of the lac genes to selected promoters in Escherichia coli using bacteriophage lambda and mu. J. Mol. Biol. 104:541–555 [DOI] [PubMed] [Google Scholar]
- 30. Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, Datsenko KA, Tomita M, Wanner BL, Mori H. 2006. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol. Syst. Biol. 2:2006.0008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Bolivar F, Rodriguez RL, Greene PJ, Betlach MC, Heyneker HL, Boyer HW, Crosa JH, Falkow S. 1977. Construction and characterization of new cloning vehicles. II. A multipurpose cloning system. Gene 2:95–113 [PubMed] [Google Scholar]
- 32. Dykxhoorn DM, St. Pierre R, Linn T. 1996. A set of compatible tac promoter expression vectors. Gene 177:133–136 [DOI] [PubMed] [Google Scholar]
- 33. Alvarez AF, Georgellis D. 2010. In vitro and in vivo analysis of the ArcB/A redox signaling pathway. Methods Enzymol. 471:205–228 [DOI] [PubMed] [Google Scholar]
- 34. Inaba K, Ito K. 2002. Paradoxical redox properties of DsbB and DsbA in the protein disulfide-introducing reaction cascade. EMBO J. 21:2646–2654 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Szajewski RP, Whitesides GM. 1980. Rate constants and equilibrium constants for thiol-disulfide interchange reactions involving oxidized glutathione. J. Am. Chem. Soc. 102:2011–2026 [Google Scholar]
- 36. Wunderlich M, Jaenicke R, Glockshuber R. 1993. The redox properties of protein disulfide isomerase (DsbA) of Escherichia coli result from a tense conformation of its oxidized form. J. Mol. Biol. 233:559–656 [DOI] [PubMed] [Google Scholar]
- 37. Miller JH. 1972. Experiments in molecular genetics, p 352–355 Cold Spring Harbor Laboratory, Cold Spring Harbor, NY [Google Scholar]
- 38. Søballe B, Poole RK. 1998. Requirement for ubiquinone downstream of cytochrome(s) b in the oxygen-terminated respiratory chains of Escherichia coli K-12 revealed using a null mutant allele of ubiCA. Microbiology 144:361–373 [DOI] [PubMed] [Google Scholar]
- 39. Wallace BJ, Young IG. 1977. Role of quinones in electron transport to oxygen and nitrate in Escherichia coli. Studies with a ubiA− menA− double quinone mutant. Biochim. Biophys. Acta 461:84–100 [DOI] [PubMed] [Google Scholar]
- 40. Zhu J, Sánchez A, Bennett GN, San K-Y. 2011. Manipulating respiratory levels in Escherichia coli for aerobic formation of reduced chemical products. Metab. Eng. 13:704–712 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Portnoy V, Scott D, Lewis NE, Tarasova Y, Osterman AL, Palsson BØ. 2010. Deletion of genes encoding cytochrome oxidases and quinol monooxygenase blocks the aerobic-anaerobic shift in Escherichia coli K-12 MG1655. Appl. Environ. Microbiol. 76:6529–6540 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Iuchi S, Chepuri V, Fu HA, Gennis RB, Lin EC. 1990. Requirement for terminal cytochromes in generation of the aerobic signal for the Arc regulatory system in Escherichia coli: study utilizing deletions and lac fusions of cyo and cyd. J. Bacteriol. 172:6020–6025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Gonidakis S, Finkel SE, Longo VD. 2011. Lifespan extension and paraquat resistance in a ubiquinone-deficient Escherichia coli mutant depend on transcription factors ArcA and TdcA. Aging 3:291–303 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Gennis RB, Stewart V. 1996. Respiration, p 217–261 In Neidhardt FC, et al. (ed), Escherichia coli and Salmonella: cellular and molecular biology, 2nd ed American Society for Microbiology, Washington, DC [Google Scholar]
- 45. Shestopalov AI, Bogachev AV, Murtazina RA, Viryasov MB, Skulachev VP. 1997. Aeration-dependent changes in composition of the quinone pool in Escherichia coli. Evidence of post-transcriptional regulation of the quinone biosynthesis. FEBS Lett. 404:272–274 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.