Skip to main content
  • Research article
  • Open access
  • Published:

Probing regulon of ArcA in Shewanella oneidensis MR-1 by integrated genomic analyses

Abstract

Background

The Arc two-component system is a global regulator controlling many genes involved in aerobic/anaerobic respiration and fermentative metabolism in Escherichia coli. Shewanella oneidensis MR-1 contains a gene encoding a putative ArcA homolog with ~81% amino acid sequence identity to the E. coli ArcA protein but not a full-length arcB gene.

Results

To understand the role of ArcA in S. oneidensis, an arcA deletion strain was constructed and subjected to both physiological characterization and microarray analysis. Compared to the wild-type MR-1, the mutant exhibited impaired aerobic growth and a defect in utilizing DMSO in the absence of O2. Microarray analyses on cells grown aerobically and anaerobically on fumarate revealed that expression of 1009 genes was significantly affected (p < 0.05) by the mutation. In contrast to E. coli ArcA, the protein appears to be dispensable in regulation of the TCA cycle in S. oneidensis. To further determine genes regulated by the Arc system, an ArcA recognition weight matrix from DNA-binding data and bioinformatics analysis was generated and used to produce an ArcA sequence affinity map. By combining both techniques, we identified an ArcA regulon of at least 50 operons, of which only 6 were found to be directly controlled by ArcA in E. coli.

Conclusion

These results indicate that the Arc system in S. oneidensis differs from that in E. coli substantially in terms of its physiological function and regulon while their binding motif are strikingly similar.

Background

Shewanella oneidensis MR-1 is a facultative gram-negative anaerobe with remarkable anaerobic respiration abilities that allow the use of a diverse array of terminal electron acceptors. These acceptors include fumarate, nitrate, nitrite, thiosulfate, elemental sulfur, trimethylamine N-oxide (TMAO), dimethyl sulfoxide (DMSO), Fe(III), Mn(III) and (IV), Cr(VI), and U(VI) [1]. Because of this exceptional metabolic flexibility and the potential use of this organism for the bioremediation of metal/radionuclide contaminants in the environment, S. oneidensis MR-1 has been extensively studied and its genome has been sequenced [2]. However, little is known on how this bacterium adopts different metabolic modes in response to the availability of oxygen. In Escherichia coli, the global regulator Fnr (fumarate nitrate regulator) plays a major role in altering gene expression between aerobic and anaerobic conditions. In contrast, S. oneidensis MR-1 appears to employ Crp (cyclic-AMP receptor protein) rather than EtrA (electron transport regulator, S. oneidensis analog to E. coli Fnr) and possibly other unidentified proteins in regulating anaerobic respiration [3–5].

Arc (aerobic respiration control) is another system playing a role in oxygen-sensing and regulating anaerobic respiration in E. coli [6]. As a classical two-component system, Arc consists of the transmembrane sensor kinase ArcB and the DNA binding response regulator ArcA [7]. Under anaerobic or microaerobic respiratory conditions, ArcB undergoes autophosphorylation by sensing the redox state of quinone pool [8, 9]. The phosphorylated ArcB then transfers a phosphate group to ArcA through a phospho-relay mechanism, resulting in phosphorylated ArcA (ArcA-P) [7, 10, 11]. ArcA-P functions as either an activator or repressor in mediating downstream genes by binding to DNA in the promoter regions of the target genes [11]. Gene expression profiling has revealed that more than a thousand genes in the E. coli genome are regulated either directly or indirectly by the ArcA protein [12, 13].

S. oneidensis MR-1 ArcA shared more than 80% in amino acid sequence identity to its homologs in a number of bacterial species in Escherichia, Salmonella, Yersinia, Erwinia, Photorhabdus, Vibrio, and Shigella [6, 8, 14–16]. In addition, the Asp54 residue in the N-terminal receiver domain and the helix-turn-helix (HTH) DNA-binding motif in the carboxy-terminal domain are structurally conserved. However, the genome lacks a definitive full-length arcB gene. ArcA of S. oneidensis MR-1 has been proven functional and involved in the oxygen response as well as in respiration of DMSO and in the detachment of cells from biofilms [14, 17, 18]. SO1327 (HptA) of S. oneidensis, exhibiting a significant degree of similarity to the Hpt domain of the E. coli ArcB, has been proposed to function to transfer phosphate groups to ArcA [14]. However, whether HptA is able to phosphorylate ArcA either in vitro or in vivo remains unanswered except that an hptA deletion strain was only slightly deficient in utilizing DMSO [14]. Furthermore, proteins equivalent to the sensor and/or the additional phosphotransfer domains remain unidentified.

The purpose of this genome-based study is to understand how the S. oneidensis MR-1 Arc system affects expression of genes under aerobic and anaerobic conditions. To this end, an arc A knockout mutational strain was constructed and subjected to physiological characterization and transcriptomic analysis. Results revealed that the mutation in arcA has a profound effect on the bacterial physiology and transcriptome. Meanwhile, an ArcA recognition weight matrix was generated using promoter regions of the core members of the Arc system to estimate the operons directly controlled by ArcA. Comparatively, this atypical Arc system differs from the E. coli Arc system substantially in terms of both functionality and regulon.

Results

Generation and verification of an arcA deletion strain

A mutagenesis system for constructing deletion mutants in S. oneidensis MR-1 has previously been developed and successfully utilized [19, 20]. The arcA deletion mutant, designated as JZ3988K (ΔarcA), was constructed using the plasmid pDS3.1 following our established procedure as described in Methods. The deletion was confirmed by PCR, DNA sequencing and phenotype complementation. The complementation plasmid pBBR-ARCA was constructed and introduced into the arcA deletion strain as described in Methods. Two consistent phenotypes were identified: aerobic growth defect and anaerobic growth defect on DMSO as presented in the next section and reported previously, respectively [14]. The wild-type and the mutant strains containing empty plasmid pBBR1MCS-5 were included as controls. In all cases, physiological differences were insignificant between the arcA mutation strain containing plasmid pBBR-ARCA and the wild-type. These results verified that the phenotype of the arcA mutant is specific to the mutation in the arcA gene.

Physiological characterization of the ΔarcA strain

Under aerobic conditions, growth of the ΔarcA strain was substantially slower than that of its parental strain MR-1 (Figure 1A). Although oxygen per se is not the direct signal for activation of the Arc system, the system functions to respond to redox conditions of growth [10]. Therefore, the observed growth difference may result from variation of culture oxygen levels in MR-1 and the mutant cultures. To rule out this possibility, levels of dissolved oxygen (DO) in both cultures were measured. As shown in Figure 1B, DO decreased quickly when cells grew up at the early stage. When cells entered the mid-log phase, DO reached the lowest point (0.06 mg/L) and remained at the level until the late stationary phase. Through the entire process, DO appeared to be irrespective of different strains but a function of cell density. All these results suggest that the Arc system of S. oneidensis has a role in the bacterial aerobiosis. Complementation of the ΔarcA strain with the pBBR-ARCA plasmid restored a growth rate similar to that of the wild-type aerobically. Under anaerobic conditions, differences in growth rate and maximum cell density between the mutant and wild-type strains were statistically insignificant when one of the following electron acceptors was used; fumarate (20 mM), nitrite (1 mM), thiosulfate (3 mM), TMAO (20 mM), MnO2 (5 mM), ferric citrate (10 mM), and cobalt(III)-EDTA (200 μM) (data not shown). In agreement with a previous report, the mutant was severely defective in utilizing DMSO (20 mM) [14].

Figure 1
figure 1

Characteristics of S. oneidensis MR-1 and mutant strains under various conditions. MR-1 (blue Diamond), JZ3988K (purple Square). (A) Growth under aerobic conditions, Complementation of JZ3988K with pBBR-ARCA (light green triangle) was also shown. (B) Dissolved oxygen concentration in MR-1 and JZ3988K cultures was plotted against OD600 values. (C) Survival rates during the stationary phase under aerobic conditions in shaken (blue diamond and purple Square) and still cultures (yellow diamond and green square) were shown.

To determine possible involvement of the Arc system in cell viability during the stationary phase in S. oneidensis, survival of the ΔarcA strain during this phase was examined in both still and shaking cultures as described in Materials and Methods. While cells of the wild-type and mutant strains died more quickly in the shaking cultures, little difference in survival rates between the wild-type and ΔarcA strains was observed under either condition. These results implicate that ArcA is dispensable in maintaining the viability of S. oneidensis cells during stasis (Figure 1C).

Global transcriptomic analysis of the ΔarcA strain

Microarray analysis was employed to dissect the transcriptomic differences elicited by the mutation in arcA during aerobiosis and anaerobiosis. For aerobiosis, the wild-type and mutant cells at exponential phase were used because the phenotypic difference was most significant during this period. Although little difference in physiology was observed between the wild-type and ΔarcA strains during anaerobiosis with fumarate as the sole electron acceptor, the exponential stage cells were collected for this study to serve two purposes. First, this may facilitate our understanding of ArcA's role during anaerobiosis. Second, this enables us to compare transcriptomes of E. coli arcA mutant to S. oneidensis arcA mutant because microarray analyses on E. coli arcA muant have been conducted under the similar conditions [12, 13]. The quality of microarray data was assessed with two approaches used as a standard in our laboratory. First, the high quality of the expression data was validated with a statistical analysis as previously described [21]. Second, 8 ORFs were selected for real-time quantitative reverse transcription-PCR (qRT-PCR) analysis with the same RNA samples used in the array hybridizations based on the level and reproducibility of changes observed in the microarray experiments. A high level of concordance (R2 = 0.96) was observed between microarray and real-time qRT-PCR data despite quantitative differences in the level of change, suggesting that the microarray results are an accurate reflection of the gene expression profile (Figure 1S in additional file 1).

In total, 1009 genes passed ANOVA statistical analysis (p < 0.05) with Benjamini and Hochberg False Discovery Rate multiple testing correction in at least one of two hybridizations between JZ3988K and MR-1, representing approximately 21.7% of the 4,648 ORFs spotted on the array (Table S1 in additional file 2). Interestingly, only 12 genes responded oppositely under aerobic and anaerobic conditions while the majority of 1009 genes responded to the arcA mutation are irrespective to the availability of oxygen. The functional class distribution of these 1009 genes is shown in Fig. 2. Genes displaying significant differences in expression due to an arcA mutation under either aerobic or anaerobic conditions were observed in almost every category. The wide distribution of putative functional roles attributed to the differentially expressed genes indicates that ArcA has a global effect on gene expression in S. oneidensis. While up to 54% of the genes showed increased expression in the absence of ArcA under aerobic conditions, the percentage of this type of genes under anaerobic conditions increased to 60%. The most noticeable differences in gene numbers between tested conditions were observed in the categories of protein synthesis (M) and nucleotide synthesis (N). Under aerobic conditions, nearly all of genes in these two categories showed reduced expression in the absence of ArcA but very few genes in these categories were affected under anaerobic conditions.

Figure 2
figure 2

Differentially expressed genes grouped by functional classification according to the TIGR S. oneidensis genome database. A, Amino acid biosynthesis; B, Biosynthesis of cofactors, prosthetic groups, and carriers; C, Cell envelope; D. Cellular processes; E, Central intermediary metabolism; F, Disrupted reading frame; G, DNA metabolism; H, Energy metabolism; I, Fatty acid and phospholipid metabolism; J, Hypothetical proteins; K, Mobile and extrachromosomal element functions; L, Protein fate; M, Protein synthesis; N, Purines, pyrimidines, nucleosides, and nucleotides; O, Regulatory functions; P, Signal transduction; Q, Transcription; R, Transport and binding proteins; S, Unclassified; T, Unknown function. Bars in black are the genes that showed decreased expression in the presence of ArcA; bars in gray are the genes that showed increased expression in the presence of ArcA.

While 1009 genes significantly affected in terms of their level of expression by the arcA mutation provide a large amount of information, it is less practical to discuss all of them in detail. Thus we generated a high-confidence list of 317 genes with at least a 2-fold change in expression and an ANOVA P value of < 0.01 (Table S2 in additional file 3). To identify co-regulated patterns of gene expression, we classified these 317 differentially expressed genes into 7 hierarchical clusters based on their log ratio of expression (Fig. 3).

Figure 3
figure 3

Hierarchical clustering of selected genes. All these genes are listed in Table S2 (in additional file 3). Expression differences (ΔarcA/MR-1) were represented by colors: red, induced, yellow, insignificant, and green, repressed. Each pattern is identified by different colors on the dendrogram and by numbers that correspond to the gene expression patterns. +O2, aerobic conditions; -O2, anaerobic conditions.

Expression Pattern I: Induced in the ΔarcA strain during either aerobiosis or anaerobiosis

Transcription of the 118 genes in this cluster was up-regulated in the ΔarcA strain under either aerobic or anaerobic conditions (Table S2 in additional file 3). 61 of these genes encode hypothetical proteins or proteins whose functions are presently unknown.

Genes encoding proteins in cellular processes include so0866 (putative minor curlin subunit CsgB), acc (aculeacin A acylase), pilU (twitching motility protein PilU), so3685 (putative curli production assembly/transport component CsgG), so3686 (putative curli production assembly/transport component CsgF), so3687 (putative curli production assembly/transport component CsgE), so4149 (putative RTX toxin), and aggA (agglutination protein). All but pilU or so4149 were reportedly involved in biofilm formation. This is not surprising because the involvement of the ArcA regulon in the development of biofilms has been firmly established in both E. coli and S. oneidensis [17, 22].

Several genes for energy metabolism were also in this cluster. Operon hoxK-hyaB-hydC encodes three subunits of the quinone-reactive Ni/Fe hydrogenase which catalyzes the reversible oxidation of molecular hydrogen and plays a central role in microbial energy metabolism [23]. The operon has been proven to be directly under the control of ArcA [14]. Correspondingly, four genes of the operon hypFBCDEA, which encode proteins required for Ni/Fe hydrogenase (encoded by operon hoxK-hyaB-hydC) maturation, were found in this cluster [24].

ArcA appears to repress expression of a number of genes encoding regulatory proteins. These included so0864 (transcriptional regulator, LuxR family), so0916 (transcriptional regulator, MarR family), rseA (sigma-E factor negative regulatory protein), so1661 (transcriptional regulator, LysR family), so1699 (transcriptional regulator), pspF (psp operon transcriptional activator), so3516 (transcriptional regulator, LacI family), so4542 (transcriptional regulator, LacI family), and rpoS (alternative sigma factor σS). Among these genes, only three have been well defined in E. coli. RseA is an anti-sigma factor that inhibits sigma E which transcribes genes that encode protein folding factors in response to extracytoplasmic stress stimuli [25]. The PspF protein belongs to the enhancer-binding protein family of sigma54-dependent activators and participates in controlling several genes involved in phage-shock, such as pspABC operon [26, 27]. In E. coli, the stationary phase alternative sigma factor σs, controls the expression of the genes involved in cell survival in response to cessation of growth (stationary phase) and provides cross-protection to various stresses [28]. Involvement of ArcA in E. coli stationary phase via catabolic control has also been established [29].

Expression Pattern II: Induced during aerobiosis but unaffected during anaerobiosis in the ΔarcA strain

Among the 317 genes, 29 showed increased expression under aerobic conditions but unchanged expression under anaerobic conditions in the mutant (Table S2 in additional file 3). It is reasonable to assume that these genes have a role during aerobiosis only.

Of the genes encoding proteins in this cluster, acnA (aconitate dydratase), aceB (malate synthetase A), and aceA (isocitrate lyase) were particularly worth noting. The acnA gene was the only one in TCA repressed by ArcA (Fig. 4). In contrast, ArcA represses transcription of the genes involved in entire TCA cycle except for acnB in E. coli [30]. Like AceB and AceA, AcnA is also a component in the glyoxylate pathway. The pathway short-circuits the TCA cycle and therefore rendering most of the TCA components unnecessary.

Figure 4
figure 4

Expression changes of genes in TCA cycle and glyoxylate pathway under aerobic conditions. Changes were recorded as the ratio of expression in ΔarcA to that in MR-1, "--" represents unaffected by the mutation, "↑" represents up-regulated in the arcA- strain.

Expression Pattern III: Induced during aerobiosis but repressed during anaerobiosis in the ΔarcA strain

Among the 317 genes, 6 showed increased expression under aerobic conditions but reduced expression under anaerobic conditions in the mutant compared to the wild-type (Table S2 in additional file 3). Most of these genes encode proteins with unknown function.

Expression Pattern IV: Unaffected during aerobiosis but induced during anaerobiosis in the ΔarcA strain

A total of 54 genes were clustered into this pattern (Table S2 in additional file 3). Among these 54 genes, 22 encode proteins in energy metabolism. The most notable observation was that all members except nuoK of the nuoA-N operon, encoding NADH dehydrogenase I, were mildly induced in the mutant only under anaerobic conditions. Studies in E. coli revealed that expression of the nuoA-N operon is repressed by ArcA-P under anaerobic conditions [31]. NADH dehydrogenase I primarily functions to couple the transfer of electrons from NADH to ubiquinone with the translocation of protons across the membrane in E. coli and Klebsiella pneumoniae [32].

The remaining 14 genes included astD (succinylglutamic semialdehyde dehydrogenase), deoC (deoxyribose-phosphate aldolase), torD (TorA specific chaperone), torA (TMAO reductase), torC (tetraheme cytochrome c), so1659 (tetraheme cytochrome c), ccoP (cytochrome c oxidase, cbb3-type, subunit III), ccoQ (cytochrome c oxidase, cbb3-type, CcoQ subunit), ccoO (cytochrome c oxidase, cbb3-type, subunit II), ccoN (cytochrome c oxidase, cbb3-type, subunit I), pflB (formate acetyltransferase), pflA (pyruvate formate-lyase 1 activating enzyme), ackA (acetate kinase), and pta (phosphate acetyltransferase). Except for so1659 and astD whose product is involved in arginine degradation, all other genes were co-regulated with the majority of their operon members [33]. The torCAD genes encode an inducible TMAO respiratory system as observed in E. coli [34, 35]. In addition to deoC, operon deoC-B-A contains other two genes deoA (thymidine phosphorylase) and deoB (phosphopentomutase) which have been classified into the functional category of purines, pyrimidines, nucleosides, and nucleotides by TIGR. The products of this operon are involved in (deoxy)ribose phosphate degradation in E. coli [36]. Cytochrome c oxidase (CcoN-O-Q-P), whose counterpart is not found in E. coli genome, has been well studied as a vital complex in oxidative phosphorylation [37]. Further exploration is needed to clarify why the operon is affected under anaerobic conditions. PflB is an enzyme catalysing the reversible reaction of pyruvate and coenzyme A into acetyl-CoA and formate after being activated by PflA under anaerobic conditions [38] while the versatile AckA-Pta have been reported to function in threonine degradation, acetate utilization, pyruvate oxidation, and mixed acid fermentation pathways under both aerobic and anaerobic conditions. AckA has been well studied as a member of ArcA regulon in E. coli [39].

Genes that belong to the category of transport and binding proteins also are enriched in this cluster. These included so1033 (putative iron-compound ABC transporter, ATP-binding protein), so1034 (iron-compound ABC transporter, permease protein), so1882 (AcrB/AcrD/AcrF family protein), modA (molybdenum ABC transporter, periplasmic molybdenum-binding protein), modC (molybdenum ABC transporter, ATP-binding protein), so4281 (putative potassium uptake protein KtrA), and ktrB (potassium uptake protein KtrB). Iron is an essential minor element for most organisms, playing vital roles in many important biological processes [40]. Although iron metabolism in E. coli is well studied, the iron-compound ABC transporter proteins in E. coli are still poorly defined. The possible E. coli genes corresponding to so1033 and so1034 are fhuA and fecC, respectively, which belong to two operons. It is not surprising that modA and modC fall into this category because the transition metal molybdenum is required for the enzymatic activities of most bacterial molybdoenzymes during anaerobiosis, including sulfite oxidase, nitrate reductase, DMSO reductase, and formate dehydrogenase [41]. KtrA and KtrB, two members of a new type of bacterial K+-uptake system, are peripheral and integral membrane proteins cooperating in K+ translocation [42]. The system appears to be widespread and functions in the adaptation of cells to hyperosmotic conditions [43, 44].

Expression Pattern V: Unaffected during aerobiosis but repressed during anaerobiosis in the ΔarcA strain

A total of 33 genes shared this expression pattern (Table S2 in additional file 3). Five genes (cymA, omcA, omcB, napB, so3980) encoding cytochrome c proteins, along with dmaA-1 and dmsB-1 encoding DMSO reductase, belong to this cluster. CymA, one of the most versatile cytochrome c proteins, supplies electrons to at least five different terminal reductases for utilizing fumarate, DMSO, nitrate, nitrite, and Fe(III) [45, 46]. OmcA and OmcB have been reported to be involved in anaerobiosis, especially in Mn(IV) reduction [47]. Genes dmaA-1 and dmsB-1 are from operon so1427-30 which is directly controlled by ArcA-P and the rest two genes encode a cytochrome c protein (SO1427) included in the cluster V and an outer membrane protein (SO1428) in this cluster [14]. While DmaA-1 and DmsB-1 are functional subunits of DMSO reductase, SO1427 and SO1428 remain uncharacterized. Two genes (so1431-2) in this cluster encoding hypothetical proteins, locating immediately after the so1427-30 operon, have been listed as members of the so1427-30 operon [14]. However, an individual operon for these two genes has been predicted by two independent studies [48, 49].

One of the most unexpected findings in this study was that three members (napD, napA, and napB) of the nap operon for nitrate reduction and so3980 (nrfA) for nitrite reduction were strongly repressed in the ΔarcA strain. In S. oneidensis, it has been demonstrated that the nap operon is essential for reduction of nitrate to nitrite in S. oneidensis [50]. Meanwhile, so3980 (nrfA) is essential for reduction of nitrite to ammonium (unpublished results). To verify this observation, expression of napA was measured by real-time qRT-PCR. The qRT-PCR results correlated well with those obtained from the microarrays as shown in Fig. 1S.

The fadA and fadB genes, consisting of an operon and encoding subunits of the fatty acid oxidation complex, belong to this cluster too. In E. coli, two fatty acid oxidation pathways (aerobic and anaerobic) have been characterized [51]. The FadAB complex functions in the aerobic fatty acid oxidation pathway only. However, it is arguable because the fadBA operon has been shown previously to be anaerobically repressed by the ArcA protein [12]. Our findings further indicated that ArcA has an effect on the fadBA operon under anaerobic conditions.

Expression Pattern VI: repressed during aerobiosis but unaffected during anaerobiosis in the ΔarcA strain

A total of 46 genes were clustered into this pattern. Most of the genes encoding ribosomal proteins and ATP synthase were found in this cluster (Table S2 in additional file 3). 20 out of 52 ribosomal structural genes (within rpl, rpm, and rps operons), along with genes (fusA-1 (translation elongation factor G), secY (SecY subunit of preprotein translocase), rimM (16S rRNA processing protein), tsf (translation elongation factor Ts), frr (ribosome recycling factor), pyrH (uridylate kinase), and radC (DNA repair protein RadC)) encoding ribosome related proteins within these three operons. 5 (atpA, atpB, atpE, atpF, and atpH) out of 9 members (atpA-I) of ATP synthase F1 and F0 were down-regulated during aerobiosis in the mutant strain. Proteins from both these subgroups belong to this macromolecule synthesis class. The ATP synthase of E. coli functions to synthesize ATP either by electron transport-link phosphorylation under aerobic conditions or by generation of an electrochemical proton gradient under anaerobic conditions and its regulation is largely under the control of the cell growth rate [52]. It is likely that the repression of these operons may result from the slower growth of mutant strain.

The remaining 13 genes included petAB (iron-sulfur & cytochrome b subunits of ubiquinol-cytochrome c reductase), lrp (leucine-responsive regulatory protein), hugA (heme transport protein), purA (adenylosuccinate synthetase) so1770-1 (glycerate kinase & GntP family permease), so3300-1 (cytochrome c proteins), and four encoding hypothetical proteins or protein with unknown functions. Cytochrome bc1 complex encoded by petA and petB contributes to the formation of membrane potential and proton gradient, which are coupled to ATP synthesis [37]. It is not surprising that expression of these two genes was consistent with expression of operon atpA-H. Lrp is a major regulatory protein involved in the expression of more than 30 operons largely in response to leucine in E. coli [53]. None of these operons have been reported to be regulated by ArcA. HugA, a predicted ATP transporter of protoheme, has been reported to be functionally related to the TonB energy transducing system [54]. Another member of the hugA operon, so3667, encoding a hypothetical protein, was also in this cluster.

Expression Pattern VII: Repressed during aerobiosis or anaerobiosis in the ΔarcA strain

This cluster contains 31 genes, 16 of which encode hypothetical proteins (Table S2 in additional file 3).

Genes in this cluster encoding transport and binding proteins include so0919 (putative serine transporter), so1821 (putative outer membrane porin), emrD (multidrug resistance protein D), so2427 (putative TonB-dependent receptor), so2865 (putative L-lysine exporter), so3099 (putative long-chain fatty acid transport protein), so3706 (NupC family protein), and so4014 (AcrB/AcrD/AcrF family protein). Genes encoding proteins in metabolic pathways include speF (inducible ornithine decarboxylase), so1427 (decaheme cytochrome c), and so3705 (putative 5-methylthioadenosine nucleosidase/S-adenosylhomocysteine nucleosidase). SpeF is the ornithine decarboxylase which helps cells against low environmental pH [55]. As the first gene of the operon so1427-30 which contains genes encoding DMSO reductase, so1427 differed in expression pattern from other members of the operon as discussed in pattern V. Also, genes so1418 (ApbE family protein), so3969 (OmpA family protein) and so4681 (glycosyl transferase, group 1 family protein) encode proteins in the functional category of cell envelope.

Screening for target operons of ArcA(-P) by EMSA

The analyses presented thus far clearly showed that S. oneidensis ArcA differs substantially from its E. coli counterpart in its physiological role and regulates a large number of genes. Although the majority of these genes may be affected indirectly, those possessing an ArcA-binding site in their upstream region are likely to be controlled directly by ArcA. In E. coli, the consensus ArcA-P binding sequence has been concluded computationally on the basis of DNA footprinting data [12]. Given the high degree of conservation in sequence and structure between ArcA proteins of E. coli and of S. oneidensis, it is reasonable to assume that ArcA of S. oneidensis binds to a sequence similar to the E. coli consensus 15-bp stretch. To verify this assumption and to facilitate the determination of the consensus ArcA(-P) binding sequence in S. oneidensis, a electrophoretic motility shift assay (EMSA) was used to examine the ArcA(-P) binding activity of upstream sequences of selected operons with purified ArcA proteins.

Expression of the S. oneidensis ArcA protein was initiated by insertion of the arc A gene into the Gateway entry vector pDONR221 using a lambda recombinase cloning strategy [56]. The entry vector containing the arcA gene was then converted to a protein expression system by recombination with the Gateway destination vector pDEST17 which resulted in the attachment of an N-terminal His-tag for protein purification. The His-tagged ArcA protein was expressed in E. coli and purified from inclusion bodies (Fig. 5A).

Figure 5
figure 5

ArcA(-P) Binding to selected promoters by EMSA. (A). Overproduced and purified recombinant S. oneidensis His6-ArcA from E. coli BL21 cells. (B). Interaction of so1661 promoter DNA with S. oneidensis His6-ArcA. The probe was prepared by PCR with SO1661-EMSA-F (33P end-labeled) and SO1661-EMSA-R primers (Table S4 in additional file 5). The EMS assay was performed with 2 nM 33P end-labeled probes and various amounts of ArcA (left panel) or ArcA-P proteins (right panel). The protein concentrations for lanes 1–9 are 0, 0.125, 0.25, 0.5, 1.0, 2.0, 4.0, 4.0, 4.0 μM, respectively. Non-specific competitor DNA, (2 μg poly dI·dC), was added (lane 8) and specific competitor (10 μM unlabeled SO1661 probe) was added (lane 9). (C). The binding assay was performed in the presence of 0, 1, or 2 μM ArcA-P and 2–5 nM radiolabeled promoter DNA 0.2 μg/μl poly(dI·dC) was used in all these binding reactions to block non-specific interactions. Promoter region of so0011 (gyrB) was included as negative control. The phosphorylation of the ArcA protein was done with carbamoyl phosphate.

In total, PCR products containing upstream intergenic regions of 30 individual operons were generated and examined for their ArcA(-P) binding activities (Table 1). Among them, gltA-sdhCAB, sucAB, icd, fdrCAB, acnB, and frdC encoding enzymes involved in the TCA cycle were chosen to validate the observation from the microarray analysis that S. oneidensis ArcA appears to be dispensable in the process [7, 57–59]. The rest of the operons selected met both or either of two criteria: a significant expression difference elicited by the arcA mutation and a 15-bp sequence in the upstream region similar to the E. coli consensus ArcA-P binding motif. In a preliminary experiment, the ArcA(-P) binding activity of the so1661 promoter region was tested with the purified unphosphorylated His-tagged S. oneidensis ArcA protein as well as the protein phosphorylated with carbamoyl phosphate in EMSA. Significant binding to the DNA probe occurred at a protein concentration of less than 0.25 μM for ArcA-P (Figure 5B), which is comparable to that from E. coli ArcA-P [7]. In contrast, the non-phoshorylated ArcA did not bind even when the protein concentration was increased to 4 μM (Figure 5B). The binding of ArcA-P to the target promoter was not reduced by addition of the nonspecific competitor poly(dI·dC) DNA, but was outcompeted by adding excess unlabeled probe (Figure 5B). These results demonstrate that phosphorylated ArcA binds the so1661 promoter in a sequence specific manner.

Table 1 Operons tested for their ArcA(-P) binding activity by EMSA

Among the tested PCR sequences in EMSA with phosphorylated ArcA, 20 showed ArcA(-P) binding activity (Figure 5C, only 8 were shown). Interestingly, all of these 20 sequences contained a 15 bp stretch sharing a high level similarity with E. coli ArcA-P binding consensus motif. Only one (so3659) such a sequence was not found to be capable of binding. In contrast, the retardation of the sequences without putative binding motif was not observed, including those (so0314, so2389, so2460) exhibiting an extremely strong expression difference between the mutant and the wild-type (data not shown). The promoter sequences of operons encoding enzymes in the TCA cycle, as expected, did not appear to interact with ArcA(-P) at all (data not shown). This result, along with our microarray data, ruled out the possible involvement of the Arc in TCA of S. oneidensis. Overall, the EMSA assays suggest that the binding motif is the most important factor determining the binding activity of promoters.

Determination of genes directly under the control of ArcA(-P) in S. oneidensis

Two recent studies suggest that more than a hundred operons are directly controlled by ArcA in E. coli [12, 13]. While our physiological and microarray analyses demonstrated that ArcA of S. oneidensis may differ substantially from the canonical Arc system in terms of both functions and regulatees, the consensus ArcA-P binding sequences from these two microorganism are alike. In other words, the binding site is more deterministic than expression changes observed in microarray analysis. To screen for operons whose upstream region contains a binding site, an ArcA-binding weight matrix was constructed with 20 PCR sequences bound by ArcA-P revealed in the EMSA (Table 1). Highly conserved stretches of 15 base pairs were found in the upstream regions of all input genes using AlignACE and a weight matrix was generated from these sequences. A sequence logo was deduced to show the frequencies scaled relative to the information content of at each position. (Figure 6) [60, 61]. Compared to the one in E. coli, subtle changes at most bases were noticeable although a high level of similarity remained. Especially, nucleotides at both ends of the binding motif are much less conserved in S. oneidensis while the 12th nucleotide (G) appears to be important for binding.

Figure 6
figure 6

Sequence logo for the ArcA-P recognition matrix in S. oneidensis. The sequences used were listed in Table 1. The sequence conservation, measured in bits, is shown as the height of a stack of letters at each base position. Sequence logo for the ArcA-P recognition matrix in E. coli was also shown as comparison. The E. coli sequences used were from the previous report by Liu and De Wulf [12].

The S. oneidensis genome was then scanned on either strand and scores of all successive 15-bp stretches were given using the log transformation method of Berg and von Hippel [62]. The average of total scores was assigned a Z score of 0 and sites with a Z score of 2.3 or greater and within 700 bp of an ORF origin were regarded as the potential ArcA-P binding sequences in S. oneidensis. The cutoff value was chosen based on an assessment that 372 genes are potential members of the ArcA regulon and the approximately 100–150 operons that may be under the transcriptional control of ArcA in E. coli [12]. In total, 209 operons containing 313 protein-encoding genes were predicted to contain potential ArcA-binding sites in their upstream regions (Table S3 in additional file 4).

By combining both microarray and weight matrix profilings (Table S1 and S3 in additional file 2 and 4, respectively), we identified at least 50 operons which could be under the direct control of ArcA in S. oneidensis (Table 2). In addition, operons next to a putative ArcA-binding site with a Z score 3.0 or above are automatically included as the candidates to be directly controlled by ArcA based on the criteria adopted in similar studies [12]. The majority of these operons encode proteins whose functions have not been determined yet, implicating a significant function shift. These newly identified operons (i.e. so1427-30, so1661, so1821) may represent lineage-specific ArcA regulon members.

Table 2 Operons that are most likely under the direct transcriptional control of ArcA

Discussion

The canonical Arc signal transduction system in γ-proteobacterial species typified by E. coli is recognized as a second global regulator that, like Fnr, mediates gene expression in response to respiratory condition changes [7, 8, 13]. In S. oneidensis, EtrA, an analog of E. coli Fnr, appears to play a negligible role in this regulatory process, leaving the Arc system the best currently recognized candidate for the role [3, 4]. In addition to high similarities in protein sequence and structure, the S. oneidensis arcA gene has been shown to be able to complement an E. coli arcA deletion mutant [14]. These findings strongly suggest a functional similarity between ArcA proteins in S. oneidensis and in other organisms and that the Asp54 residue is the phosphorylation site if required. A homolog to the arcB gene in S. oneidensis MR-1 has yet to be identified in the genome. Further comparative analyses of multiple Alteromonadaceae genomes indicate that this type of Arc system is in fact common among the Alteromonadaceae. Unfortunately, the 'atyptical' Arc systems have been largely overlooked possibly because of the sequence and structure conservation of ArcAs.

In this study, we have attempted to understand the major physiological changes mediated by ArcA and define its regulon with a comparison with ArcA of the canonical Arc system in E. coli. The two systems differ significantly from each other in several key aspects. Firstly, physiologically, one of major questions about the Arc system is whether it regulates any aspect of aerobic respiration. Unexpectedly, S. oneidensis ArcA is directly involved in aerobic metabolism. In E. coli, it is believed that the Arc system regulates gene expression in response to anaerobic conditions under which ArcB phosphorylates ArcA [8]. Correspondingly, the maximum growth rate of an E. coli arcA mutant was not significantly different from that of the wild-type when grown on a variety of media if oxygen deprivation is excluded [22, 30, 63–65]. Similarly, the absence of ArcA did not show any effect on cell morphology and growth characteristics of Salmonella enterica serovar Enteritidis under aerobic conditions [15]. All these results indicate that ArcA has a very limited role, if any, in aerobic respiration in these bacteria. Secondly, ArcA of S. oneidensis appears to be irrelevant to survival during stationary phase. The E. coli ArcA is heavily involved in starvation-induced modulations of gene expression and therefore plays a key role in the bacterial stasis survival [29, 66, 67]. Thirdly, E. coli ArcA proteins directly control the TCA cycle while the S. oneidensis ArcA controls DMSO reduction directly [14]. All these differences suggest that the S. oneidensis ArcA functionally deviates from the canonical one considerably.

In spite of significant difference in their physiological roles, the activation mechanism by phosphorylation and the target sequences of ArcA proteins of E. coli and S. oneidensis share a high level of similarity. In this study, the EMSA results reinstated that a binding motif in the promoter region rather than expression differences of target genes appears to be more crucial for binding. Nevertheless, a combination of binding motif and expression difference promotes more accurate prediction. With the combination, up to 50 S. oneidensis operons are identified while at least 82 operons are reportedly under direct control of ArcA in E. coli [12]. Given that up to 2183 genes are predicted to be in common in E. coli (~48.8%) and S. oneidensis (~51.4%) genomes [68], the number of overlapping operons (6) is surprisingly small (Figure 7). The majority of members in the E. coli ArcA regulon involved in metabolism are not identified in the S. oneidensis ArcA regulon, implicating that a significant difference in ArcA regulons of these two organisms has evolved. It is worth noting that none of the operons encoding TCA enzymes were located next to these high-confident ArcA-binding sites, consistent with our observation that these TCA operons were not affected by the arcA mutation. In E. coli, on the contrary, the promoter region of most TCA genes contains an ArcA-binding site with z score above 3.0 [12]. It is possible that ArcA-independent expression of S. oneidensis TCA genes may be largely due to the loss of ArcA-binding sites in their promoter regions.

Figure 7
figure 7

Common genes in S. oneidensis and E. coli genomes and in ArcA regulons of these two bacteria. Common genes in S. oneidensis and E. coli genomes are obtained from [68] using default similarity cutoffs (maximum E-value, 1e-5; minimum percent identity, 30).

Similarly, S. oneidensis ArcA may acquire controls over new genes once an ArcA-binding site emerged in their promoter regions through evolution. Genes (so1427-30) for DMSO reductase and related proteins serve as a good example. DMSO reduction pathway shares most of components for anaerobic respiration on all other electron acceptors except TMAO [46]. In E. coli, the Arc system functions as a global regulator of respiratory gene expression under microaerobic and anaerobic growth conditions [8–10]. As a result, utilization of many compounds anaerobically is found to be affected simultaneously. In contrast, the arcA mutant is defective in DMSO respiration only, making it hard to accept that S. oneidensis ArcA functions in a similar way. Given that a predicted ArcA-binding site is found within so1427 promoter region, we propose that this operon become a new number of ArcA regulon only because the binding site shows up, presumably by chance.

Conclusion

This study provides the first comprehensive profile to elucidate the functions of the atypical Arc system in S. oneidensis, compared to the canonical one in E. coli. Our findings demonstrated that two Arc systems are significantly different from each other with respect to the physiological functions and the regulons although the sequences and binding motif are highly similar. S. oneidensis ArcA does not appear to be involved in regulation of TCA cycle while ArcA in E. coli repress the genes involved in the TCA cycle under anaerobic condition. More than 50 operons were confidently identified as members of the Shewanella ArcA regulon, but a much larger number of members are expected. However, only a very limited number of the regulon members are shared by the E. coli ArcA regulon. The significant differences in both physiology of arcA mutants and regulon of ArcA proteins of these two microorganisms may simply be due to the differences in lifestyle, metabolism, and gene content between them. Further molecular characterization of the lineage-specific ArcA regulon members identified in this study is needed to dissect the functional diversity and ultimately the evolution of the Arc system in γ-proteobacteria.

Methods

Bacterial strains, plasmids, and culture conditions

A list of all bacterial strains and plasmids used in this study is given in Table 3. E. coli and S. oneidensis strains under aerobic conditions were grown in Luria-Bertani (LB, Difco, Detroit, MI) medium at 37°C and room temperature for genetic manipulation, respectively. When needed, the growth medium was supplemented with antibiotics at the following concentrations: ampicillin at 50 μg/ml and gentamycin at 15 μg/ml. The suicide vector pDS3.1 has been described elsewhere [20].

Table 3 Strains and plasmids used in this study

Disruption of arcA and complementation of the resulting arcA mutant

An arcA deletion strain was constructed. Primers used for generating PCR products for mutagenesis are listed in Table S4 (Table S4 in additional file 5). In brief, two fragments flanking arcA were amplified by PCR with primers SO3988-5-F and SO3988-5-R, primers SO3988-3-F and SO3988-3-R, respectively, and purified using the QIAquick PCR purification kit (Qiagen, Chatsworth, CA). Fusion PCR products generated using the amplified fragments as templates with primers SO3988-5-F and SO3988-3-R as described elsewhere [69]. The resulting fusion fragment was ligated into the Xcm I site of plasmid pDS3.1 and the resulting mutagenesis vectors (pDS-ARCAK) were transformed into the plasmid donor strain, E. coli WM3064 [70]. Plasmids pDS-ARCAK in WM3064, grown on LB supplemented with 0.3 mM diaminopimelic acid (DAP), were further transferred to MR-1 by conjugation [20]. Integration of the mutagenesis construct into the chromosome was selected by gentamycin resistance and confirmed by PCR amplification. Verified transconjugants were grown in LB broth in the absence of NaCl and plated on LB supplemented with 10% of sucrose. Gentamycin-sensitive and sucrose-resistant colonies were screened by PCR for the deletion of arcA. The deletion mutation was then verified by sequencing of the mutated region, and the deletion strain was designated as JZ3988K (ΔarcA).

For complementation, a 1.4-kb DNA fragment containing arcA and its native promoter was generated by PCR amplification with MR-1 genomic DNA as the template using primers SO3988-COM-F and SO3988-COM-R as listed in Table S4 (Table S4 in additional file 5). This fragment was digested with Sac I (underlined) and ligated to Sac I-digested pBBR1MCS-5 to form pBBR-ARCA [71], which was electroporated into WM3064. Introduction of pBBR-ARCA into JZ3988K was done by mating with WM3064 hosting pBBR-ARCA, and gentamycin-resistant colonies were selected. The presence of pBBR-ARCA in JZ3988K was confirmed by plasmid purification and restriction enzyme digestion.

Physiological characterization of the mutation strain under various conditions

M1 defined medium containing 0.02% (w/v) of vitamin-free Casamino Acids and 15 mM lactate was used in all physiological experiments [72]. Growth of the deletion strain under aerobic or anaerobic conditions was determined by recording growth curves in triplicate with a Bioscreen C microbiology reader (Labsystems Oy, Helsinki, Finland) with MR-1 as the control. For aerobic growth, exponential phase cultures were diluted to approximately ~1 × 105 cells/ml in fresh medium, and 400 μl was transferred to the honeycomb plate wells of the Bioscreen C reader. The cultures were shaken at medium intensity continuously, and the turbidity was measured every 30 min at 600 nm and DO (dissolved oxygen) was recorded every hour with an Accumet XL40 meter (Fisher Scientific). For anaerobic growth, exponential phase cultures grown aerobically were centrifuged, purged in nitrogen and suspended in fresh medium to approximately ~1 × 105 cells/ml in an anaerobic glove box. Electron acceptors tested in this study included fumarate (20 mM), nitrate (2 mM), nitrite (1 mM), thiosulfate (3 mM), TMAO (20 mM), and DMSO (20 mM). For electron acceptors containing metals including MnO2 (5 mM), ferric citrate (10 mM), and cobalt(III)-EDTA (200 μM), growth was monitored by the color change of the cultures and cell counting under a microscope (Nikon Optiphot, Nikon, Japan).

Survival of MR-1 and the ΔarcA strain during the stationary phase was examined. Cultures were grown from a single colony under aerobic conditions with vigorous shaking. After the onset of stationary phase, the cultures were divided into two parts. One was kept in the incubator with vigorous shaking and the other was kept still. The cultures were serially diluted into LB and plated onto LB plates every 12 h. Plates from dilutions that gave 100 to 250 colony form units (CFU) per plate were used to minimize statistical variation due to small sample sizes. Experiments were done in triplicate.

Microarray analysis

For each strain under aerobic conditions, 100 ml of M1 medium in a 500 ml shake flask was inoculated with fresh overnight culture to OD600 of 0.01 and then was divided into four aliquots (biological replicates) and shaken on a rotary platform (250 rpm) until mid-log phase (OD ≈ 0.4 at 600 nm). For anaerobic cultures, 500 ml of M1 supplemented with 10 mM fumarate as the electron acceptor was inoculated to an OD600 of 0.01 and then divided into four aliquots (biological replicates) and kept in an anaerobic chamber until mid-log phase (OD ≈ 0.15 at 600 nm). All cultures were centrifuged at 8000 rpm in a Sorvall RC5C plus for 3 min at the room temperature and the pellet was frozen immediately in liquid nitrogen and stored at -80°C.

DNA microarrays were constructed using PCR-amplified fragments of each annotated open reading frame from S. oneidensis MR-1, as previously described [21]. Total RNA extraction, cDNA labeling, hybridization, and slide scanning were conducted according to the standard procedure used in our lab [21, 73, 74]. LOWESS was used to normalize the data set which subsequently was subjected to statistical analysis by analysis of variance (ANOVA) with Benjamini and Hochberg False Discovery Rate as multiple testing correction. Genes with at an ANOVA P value of < 0.05 were considered significantly differentially expressed and were listed in supplemental materials. For discussion in text, a list of genes with at least a 2-fold change in expression and an ANOVA P value of < 0.01 were used. All raw data (MIAME) have been deposited to NCBI GEO (GSE7973).

Real-time quantitative RT-PCR (qRT-PCR)

qRT-PCR was performed with primers listed in Table S4 (Table S4 in additional file 5) as described previously [21, 74]. A 100-bp fragment of the acnA gene, which was amplified by PCR with genomic DNA as the template, was used to construct the standard curve. The expression of each gene was determined from three replicates on a single real-time qRT-PCR experiment. The expression ratio was recorded as the fold difference in quantity of real-time qRT-PCR product from samples grown at the treatment vs. control.

Expression and purification of S. oneidensis ArcA protein

Plasmids pDONR221 and pDEST17 and E. coli BL21 (DE3) Star cells were obtained from Invitrogen. To create pDEST17-ArcA, the ArcA encoding ORF was first cloned into pDONR221 by using ArcA-up/down primers (Table S4 in additional file 5), and then transferred into pDEST17 for protein expression by Gateway recombination reactions. All of these plasmid constructs were verified by DNA sequencing. The expression of ArcA in E. coli BL21(DE3) Star cells was induced with 0.5 mM IPTG from mid-log phase (OD600 = 0.5–0.6) at 30°C. The cells were grown to saturation and then collected by centrifugation, resuspended in lysis buffer (50 mM Tris/HCl, pH 7.5, 200 mM NaCl, 1 mM MgCl2, 10 mM β-mercaptoethanol, 1 mM PMSF, 5 μg/mL DNaseI), and broken by passage twice through a French press (10,000 psi). The resulting inclusion body pellets were solubilized with 20 mM Tris/HCl (pH 8.0), 5 M urea and 100 mM NaCl, and the ArcA protein was further purified by using Talon resin columns (BD Biosciences®) under denaturing conditions according to manufacturer instructions. To renature the protein, the eluted fractions containing ArcA protein were collected, diluted into 0.8 M urea, 20 mM Tris/HCl (pH 8.0), 1 mM EDTA by sequential dilutions, and then dialyzed against 20 mM Tris/HCl (pH 7.5). Finally the ArcA protein was concentrated to ~0.6 mg/ml.

Phosphorylation of ArcA and electrophoretic motility shift assay (EMSA)

Phosphorylation of purified ArcA protein was performed in buffer containing 100 mM Tris/HCl (pH 7.0), 10 mM MgCl2, 125 mM KCl, 50 mM dilithium carbamoyl phosphate for 60 minutes at room temperature as described [7]. The probes used for EMSA were prepared by PCR with 33P end-labeled primers (Table S4 in additional file 5). The binding reaction was performed with ~25–50 fmol (~2–5 nM) labeled probes and various amount of protein in 12 μl binding buffer containing 100 mM Tris/HCl (pH 7.4), 20 mM KCl, 10 mM MgCl2, 2 mM DTT, 0.2 μg/μl poly(dI·dC), and 10% glycerol at 15°C for 60 minutes and resolved on pre-run 4.8% polyacrylamide native gels [7]. The band shifts were visualized by autoradiography.

ArcA weight matrix development and identification of putative ArcA-binding sites

AlignACE was used to screen for a common ArcA-binding motif within promoter regions of ArcA-controlled operons predicted either by DNA footprinting or microarray analysis [58]. The identified ArcA-binding motifs of 15 bp were transformed to a weight matrix using the method of Berg and von Hippel [62]. The whole genome was then scaned for putative ArcA-binding motifs with the weight matrix.

References

  1. Nealson KH, Scott J: Ecophysiology of the genus Shewanella. The Prokaryotes. Edited by: Dworkin M. 2004, Springer-Verlag

    Google Scholar 

  2. Heidelberg JF, Paulsen IT, Nelson KE, Gaidos EJ, Nelson WC, Read TD, Eisen JA, Seshadri R, Ward N, Methe B, Clayton RA, Meyer T, Tsapin A, Scott J, Beanan M, Brinkac L, Daugherty S, DeBoy RT, Dodson RJ, Durkin AS, Haft DH, Kolonay JF, Madupu R, Peterson JD, Umayam LA, White O, Wolf AM, Vamathevan J, Weidman J, Impraim M, Lee K, Berry K, Lee C, Mueller J, Khouri H, Gill J, Utterback TR, McDonald LA, Feldblyum TV, Smith HO, Venter JC, Nealson KH, Fraser CM: Genome sequence of the dissimilatory metal ion-reducing bacterium Shewanella oneidensis. Nat Biotech. 2002, 20 (11): 1118-1123. 10.1038/nbt749.

    Article  Google Scholar 

  3. Beliaev AS, Thompson DK, Fields MW, Wu L, Lies DP, Nealson KH, Zhou J: Microarray transcription profiling of a Shewanella oneidensis etrA mutant. J Bacteriol. 2002, 184 (16): 4612-4616. 10.1128/JB.184.16.4612-4616.2002.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Maier TM, Myers CR: Isolation and characterization of a Shewanella putrefaciens MR-1 electron transport regulator etrA mutant: Reassessment of the role of EtrA. J Bacteriol. 2001, 183 (16): 4918-4926. 10.1128/JB.183.16.4918-4926.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Saffarini DA, Schultz R, Beliaev A: Involvement of cyclic AMP (cAMP) and cAMP receptor protein in anaerobic respiration of Shewanella oneidensis. J Bacteriol. 2003, 185 (12): 3668-3671. 10.1128/JB.185.12.3668-3671.2003.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Iuchi S, Lin ECC: arcA (dye), a global regulatory gene in Escherichia coli mediating repression of enzymes in aerobic pathways. PNAS. 1988, 85 (6): 1888-1892. 10.1073/pnas.85.6.1888.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Lynch AS, Lin EC: Transcriptional control mediated by the ArcA two-component response regulator protein of Escherichia coli: characterization of DNA binding at target promoters. J Bacteriol. 1996, 178 (21): 6238-6249.

    PubMed  PubMed Central  Google Scholar 

  8. Georgellis D, Kwon O, Lin ECC: Quinones as the redox signal for the Arc two-component system of bacteria. Science. 2001, 292 (5525): 2314-2316. 10.1126/science.1059361.

    Article  PubMed  Google Scholar 

  9. Malpica R, Franco B, Rodriguez C, Kwon O, Georgellis D: Identification of a quinone-sensitive redox switch in the ArcB sensor kinase. PNAS. 2004, 101 (36): 13318-13323. 10.1073/pnas.0403064101.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Georgellis D, Lynch AS, Lin EC: In vitro phosphorylation study of the arc two-component signal transduction system of Escherichia coli. J Bacteriol. 1997, 179 (17): 5429-5435.

    PubMed  PubMed Central  Google Scholar 

  11. Kwon O, Georgellis D, Lin ECC: Phosphorelay as the sole physiological route of signal transmission by the Arc two-component system of Escherichia coli. J Bacteriol. 2000, 182 (13): 3858-3862. 10.1128/JB.182.13.3858-3862.2000.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Liu X, De Wulf P: Probing the ArcA-P modulon of Escherichia coli by whole genome transcriptional analysis and sequence recognition profiling. J Biol Chem. 2004, 279 (13): 12588-12597. 10.1074/jbc.M313454200.

    Article  PubMed  Google Scholar 

  13. Salmon KA, Hung S-p, Steffen NR, Krupp R, Baldi P, Hatfield GW, Gunsalus RP: Global gene expression profiling in Escherichia coli K12: EFFECTS OF OXYGEN AVAILABILITY AND ArcA. J Biol Chem. 2005, 280 (15): 15084-15096. 10.1074/jbc.M414030200.

    Article  PubMed  Google Scholar 

  14. Gralnick JA, Brown CT, Newman DK: Anaerobic regulation by an atypical Arc system in Shewanella oneidensis. Mol Microbiol. 2005, 56 (5): 1347-1357. 10.1111/j.1365-2958.2005.04628.x.

    Article  PubMed  Google Scholar 

  15. Lu S, Killoran PB, Fang FC, Riley LW: The global regulator ArcA controls resistance to reactive nitrogen and oxygen intermediates in Salmonella enterica Serovar Enteritidis. Infect Immun. 2002, 70 (2): 451-461. 10.1128/IAI.70.2.451-461.2002.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Sengupta N, Paul K, Chowdhury R: The global regulator ArcA modulates expression of virulence factors in Vibrio cholerae. Infect Immun. 2003, 71 (10): 5583-5589. 10.1128/IAI.71.10.5583-5589.2003.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Thormann KM, Saville RM, Shukla S, Spormann AM: Induction of rapid detachment in Shewanella oneidensis MR-1 biofilms. J Bacteriol. 2005, 187: 1014-1021. 10.1128/JB.187.3.1014-1021.2005.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Vanrobaeys F, Devreese B, Lecocq E, Rychlewski L, De Smet L, Van Beeumen J: Proteomics of the dissimilatory iron-reducing bacterium Shewanella oneidensis MR-1, using a matrix-assisted laser desorption/ionization-tandem-time of flight mass spectrometer. Proteomics. 2003, 3: 2249-2257. 10.1002/pmic.200300476.

    Article  PubMed  Google Scholar 

  19. Gao W, Liu Y, Giometti C, Tollaksen S, Khare T, Wu L, Klingeman D, Fields M, Zhou J: Knock-out of SO1377 gene, which encodes the member of a conserved hypothetical bacterial protein family COG2268, results in alteration of iron metabolism, increased spontaneous mutation and hydrogen peroxide sensitivity in Shewanella oneidensis MR-1. BMC Genomics. 2006, 7 (1): 76-10.1186/1471-2164-7-76.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Wan X-F, VerBerkmoes NC, McCue LA, Stanek D, Connelly H, Hauser LJ, Wu L, Liu X, Yan T, Leaphart A, Hettich RL, Zhou J, Thompson DK: Transcriptomic and proteomic characterization of the Fur modulon in the metal-reducing bacterium Shewanella oneidensis. J Bacteriol. 2004, 186 (24): 8385-8400. 10.1128/JB.186.24.8385-8400.2004.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Gao H, Wang Y, Liu X, Yan T, Wu L, Alm E, Arkin A, Thompson DK, Zhou J: Global transcriptome analysis of the heat shock response of Shewanella oneidensis. J Bacteriol. 2004, 186 (22): 7796-7803. 10.1128/JB.186.22.7796-7803.2004.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Junker LM, Peters JE, Hay AG: Global analysis of candidate genes important for fitness in a competitive biofilm using DNA-array-based transposon mapping. Microbiology. 2006, 152 (8): 2233-2245. 10.1099/mic.0.28767-0.

    Article  PubMed  Google Scholar 

  23. Vignais PM, Colbeau A: Molecular biology of microbial hydrogenases. Curr Issues Mol Biol. 2004, 6: 159-188.

    PubMed  Google Scholar 

  24. Hube M, Blokesch M, Bock A: Network of hydrogenase maturation in Escherichia coli: Role of accessory proteins HypA and HybF. J Bacteriol. 2002, 184 (14): 3879-3885. 10.1128/JB.184.14.3879-3885.2002.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Tam C, Collinet B, Lau G, Raina S, Missiakas D: Interaction of the conserved region 4.2 of sigma E with the RseA anti-sigma factor. J Biol Chem. 2002, 277 (30): 27282-27287. 10.1074/jbc.M202881200.

    Article  PubMed  Google Scholar 

  26. Jovanovic G, Dworkin J, Model P: Autogenous control of PspF, a constitutively active enhancer-binding protein of Escherichia coli. J Bacteriol. 1997, 179 (16): 5232-5237.

    PubMed  PubMed Central  Google Scholar 

  27. Jovanovic G, Lloyd LJ, Stumpf MPH, Mayhew AJ, Buck M: Induction and function of the phage shock protein extracytoplasmic stress response in Escherichia coli. J Biol Chem. 2006, 281 (30): 21147-21161. 10.1074/jbc.M602323200.

    Article  PubMed  Google Scholar 

  28. Nystrom T: STATIONARY-PHASE PHYSIOLOGY. Annu Rev Microbiol. 2004, 58 (1): 161-181. 10.1146/annurev.micro.58.030603.123818.

    Article  PubMed  Google Scholar 

  29. Nystrom T, Larsson C, Gustafsson L: Bacterial defense against aging: Role of the Escherichia coli ArcA regulator in gene expression, readjusted energy flux and survival during stasis. Embo Journal. 1996, 15 (13): 3219-3228.

    PubMed  PubMed Central  Google Scholar 

  30. Perrenoud A, Sauer U: Impact of global transcriptional regulation by ArcA, ArcB, Cra, Crp, Cya, Fnr, and Mlc on glucose catabolism in Escherichia coli. J Bacteriol. 2005, 187 (9): 3171-3179. 10.1128/JB.187.9.3171-3179.2005.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Bongaerts J, Zoske S, Weidner U, Linden G: Transcriptional regulation of the proton translocating NADH dehydrogenase (nuoA-N) of Escherichia coli by electron acceptors, electron donors and gene regulators. Mol Microbiol. 1995, 16 (3): 521-534. 10.1111/j.1365-2958.1995.tb02416.x.

    Article  PubMed  Google Scholar 

  32. Stolpe S, Friedrich T: The Escherichia coli NADH:ubiquinone oxidoreductase (complex I) is a primary proton-pump but may be capable of secondary sodium antiport. J Biol Chem. 2004, M311242200

    Google Scholar 

  33. Schneider BL, Kiupakis AK, Reitzer LJ: Arginine catabolism and the arginine succinyltransferase pathway in Escherichia coli. J Bacteriol. 1998, 180 (16): 4278-4286.

    PubMed  PubMed Central  Google Scholar 

  34. Bordi C, Ansaldi M, Gon S, Jourlin-Castelli C, Iobbi-Nivol C, Mejean V: Genes regulated by TorR, the trimethylamine oxide response regulator of Shewanella oneidensis. J Bacteriol. 2004, 186 (14): 4502-4509. 10.1128/JB.186.14.4502-4509.2004.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Mejean V, lobbi-Nivol C, Lepelletier M, Giordano G, Chippaux M, Pascal M-C: TMAO anaerobic respiration in Escherichia coli: involvement of the tor operon. Mol Microbiol. 1994, 11 (6): 1169-1179. 10.1111/j.1365-2958.1994.tb00393.x.

    Article  PubMed  Google Scholar 

  36. Neuhard J, Kelln RA: Biosynthesis and conversions of pyrimidines. Escherichia coli and Salmonella. Edited by: Neidhardt FC and others. 1996, Washington, DC: American Society for Microbiology, 580-599.

    Google Scholar 

  37. Thony-Meyer L: Biogenesis of respiratory cytochromes in bacteria. Microbiol Mol Biol Rev. 1997, 61 (3): 337-376.

    PubMed  PubMed Central  Google Scholar 

  38. Korsa I, Bock A: Characterization of fhlA mutations resulting in ligand-independent transcriptional activation and ATP hydrolysis. J Bacteriol. 1997, 179: 41-45.

    PubMed  PubMed Central  Google Scholar 

  39. Shalel-Levanon S, San K-Y, Bennett GN: Effect of ArcA and FNR on the expression of genes related to the oxygen regulation and the glycolysis pathway Escherichia coli under microaerobic growth conditions. Biotechnol Bioeng. 2005, 92 (2): 147-159. 10.1002/bit.20583.

    Article  PubMed  Google Scholar 

  40. Andrews SC, Robinson AK, Rodriguez-Quinones F: Bacterial iron homeostasis. FEMS Microbiol Rev. 2003, 27 (2–3): 215-237. 10.1016/S0168-6445(03)00055-X.

    Article  PubMed  Google Scholar 

  41. Kisker C, Schindelin aH, Rees DC: MOLYBDENUM-COFACTOR-CONTAINING ENZYMES:Structure and Mechanism. Annu Rev Biochem. 1997, 66 (1): 233-267. 10.1146/annurev.biochem.66.1.233.

    Article  PubMed  Google Scholar 

  42. Nakamura T, Yuda R, Unemoto T, Bakker EP: KtrAB, a new type of bacterial K+-uptake system from Vibrio alginolyticus. J Bacteriol. 1998, 180 (13): 3491-3494.

    PubMed  PubMed Central  Google Scholar 

  43. Holtmann G, Bakker EP, Uozumi N, Bremer E: KtrAB and KtrCD: Two K+ uptake systems in Bacillus subtilis and their role in adaptation to hypertonicity. J Bacteriol. 2003, 185 (4): 1289-1298. 10.1128/JB.185.4.1289-1298.2003.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Matsuda N, Kobayashi H, Katoh H, Ogawa T, Futatsugi L, Nakamura T, Bakker EP, Uozumi N: Na+-dependent K+ uptake Ktr system from the cyanobacterium synechocystis sp. PCC 6803 and its role in the early phases of cell adaptation to hyperosmotic shock. J Biol Chem. 2004, M407268200

    Google Scholar 

  45. Myers JM, Myers CR: Role of the tetraheme cytochrome CymA in anaerobic electron transport in cells of Shewanella putrefaciens MR-1 with normal levels of menaquinone. J Bacteriol. 2000, 182 (1): 67-75.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Schwalb C, Chapman SK, Reid GA: The tetraheme cytochrome CymA is required for anaerobic respiration with dimethyl sulfoxide and nitrite in Shewanella oneidensis. Biochemistry. 2003, 42 (31): 9491-9497. 10.1021/bi034456f.

    Article  PubMed  Google Scholar 

  47. Myers JM, Myers CR: Role for outer membrane cytochromes OmcA and OmcB of Shewanella putrefaciens MR-1 in reduction of manganese dioxide. Appl Environ Microbiol. 2001, 67 (1): 260-269. 10.1128/AEM.67.1.260-269.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  48. The Virtual Institute for Microbial Stress and Survival Database. [http://www.microbesonline.org]

  49. The Pathway/Genome Database. [http://biocyc.org/SONE211586/]

  50. Cruz-Garcia C, Murray AE, Klappenbach JA, Stewart V, Tiedje JM: Respiratory nitrate ammonification by Shewanella oneidensis MR-1. J Bacteriol. 2007, 189 (2): 656-662. 10.1128/JB.01194-06.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Campbell JW, Morgan-Kiss RM, E Cronan J: A new Escherichia coli metabolic competency: growth on fatty acids by a novel anaerobic beta-oxidation pathway. Mol Microbiol. 2003, 47 (3): 793-805. 10.1046/j.1365-2958.2003.03341.x.

    Article  PubMed  Google Scholar 

  52. Kasimoglu E, Park SJ, Malek J, Tseng CP, Gunsalus RP: Transcriptional regulation of the proton-translocating ATPase (atpIBEFHAGDC) operon of Escherichia coli: control by cell growth rate. J Bacteriol. 1996, 178 (19): 5563-5567.

    PubMed  PubMed Central  Google Scholar 

  53. Calvo JM, Matthews RG: The leucine-responsive regulatory protein, a global regulator of metabolism in Escherichia coli. Microbiol Mol Biol Rev. 1994, 58 (3): 466-490.

    Google Scholar 

  54. Henderson DP, Wyckoff EE, Rashidi CE, Verlei H, Oldham AL: Characterization of the Plesiomonas shigelloides genes encoding the heme iron utilization system. J Bacteriol. 2001, 183 (9): 2715-2723. 10.1128/JB.183.9.2715-2723.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Kashiwagi K, Suzuki T, Suzuki F, Furuchi T, Kobayashi H, Igarashi K: Coexistence of the genes for putrescine transport protein and ornithine decarboxylase at 16 min on Escherichia coli chromosome. J Biol Chem. 1991, 266 (31): 20922-20927.

    PubMed  Google Scholar 

  56. Hartley JL, Temple GF, Brasch MA: DNA Cloning Using In Vitro Site-Specific Recombination. Genome Res. 2000, 10 (11): 1788-1795. 10.1101/gr.143000.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Chao G, Shen J, Tseng CP, Park SJ, Gunsalus RP: Aerobic regulation of isocitrate dehydrogenase gene (icd) expression in Escherichia coli by the arcA and fnr gene products. J Bacteriol. 1997, 179 (13): 4299-4304.

    PubMed  PubMed Central  Google Scholar 

  58. Park SJ, Gunsalus RP: Oxygen, iron, carbon, and superoxide control of the fumarase fumA and fumC genes of Escherichia coli: role of the arcA, fnr, and soxR gene products. J Bacteriol. 1995, 177 (21): 6255-6262.

    PubMed  PubMed Central  Google Scholar 

  59. Park SJ, McCabe J, Turna J, Gunsalus RP: Regulation of the citrate synthase (gltA) gene of Escherichia coli in response to anaerobiosis and carbon supply: role of the arcA gene product. J Bacteriol. 1994, 176 (16): 5086-5092.

    PubMed  PubMed Central  Google Scholar 

  60. Roth FP, Hughes JD, Estep PW, Church GM: Finding DNA regulatory motifs within unaligned noncoding sequences clustered by whole-genome mRNA quantitation. Nat Biotech. 1998, 16 (10): 939-945. 10.1038/nbt1098-939.

    Article  Google Scholar 

  61. Schneider TD, Stephens RM: Sequence logos: a new way to display consensus sequences. Nucl Acids Res. 1990, 18 (20): 6097-6100. 10.1093/nar/18.20.6097.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Berg OG, von Hippel PH: Selection of DNA binding sites by regulatory proteins: Statistical-mechanical theory and application to operators and promoters. J Mol Biol. 1987, 193 (4): 723-743. 10.1016/0022-2836(87)90354-8.

    Article  PubMed  Google Scholar 

  63. Alexeeva S, de Kort B, Sawers G, Hellingwerf KJ, de Mattos MJT: Effects of limited aeration and of the ArcAB system on intermediary pyruvate catabolism in Escherichia coli. J Bacteriol. 2000, 182 (17): 4934-4940. 10.1128/JB.182.17.4934-4940.2000.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Alexeeva S, Hellingwerf KJ, Teixeira de Mattos MJ: Quantitative assessment of oxygen availability: Perceived aerobiosis and its effect on flux distribution in the respiratory chain of Escherichia coli. J Bacteriol. 2002, 184 (5): 1402-1406.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Alexeeva S, Hellingwerf KJ, Teixeira de Mattos MJ: Requirement of ArcA for redox regulation in Escherichia coli under microaerobic but not anaerobic or aerobic conditions. J Bacteriol. 2003, 185 (1): 204-209. 10.1128/JB.185.1.204-209.2003.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Chang D-E, Smalley DJ, Conway T: Gene expression profiling of Escherichia coli growth transitions: an expanded stringent response model. Mol Microbiol. 2002, 45 (2): 289-306. 10.1046/j.1365-2958.2002.03001.x.

    Article  PubMed  Google Scholar 

  67. Mika F, Hengge R: A two-component phosphotransfer network involving ArcB, ArcA, and RssB coordinates synthesis and proteolysis of sigma S (RpoS) in E. coli. Genes Dev. 2005, 19 (22): 2770-2781. 10.1101/gad.353705.

    Article  PubMed  PubMed Central  Google Scholar 

  68. The Integrated Microbial Genomes. [http://img.jgi.doe.gov]

  69. Link AJ, Phillips D, Church GM: Methods for generating precise deletions and insertions in the genome of wild-type Escherichia coli: application to open reading frame characterization. J Bacteriol. 1997, 179 (20): 6228-6237.

    PubMed  PubMed Central  Google Scholar 

  70. Saltikov CW, Newman DK: Genetic identification of a respiratory arsenate reductase. PNAS. 2003, 100 (19): 10983-10988. 10.1073/pnas.1834303100.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Kovach ME, Elzer PH, Steven Hill D, Robertson GT, Farris MA, Roop RM, Peterson KM: Four new derivatives of the broad-host-range cloning vector pBBR1MCS, carrying different antibiotic-resistance cassettes. Gene. 1995, 166 (1): 175-176. 10.1016/0378-1119(95)00584-1.

    Article  PubMed  Google Scholar 

  72. Myers CR, Myers JM: MtrB is required for proper incorporation of the cytochromes OmcA and OmcB into the outer membrane of Shewanella putrefaciens MR-1. Appl Environ Microbiol. 2002, 68 (11): 5585-5594.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Gao H, Yang ZK, Wu L, Thompson DK, Zhou J: Global transcriptome analysis of the cold shock response of Shewanella oneidensis MR-1 and mutational analysis of its classical cold shock proteins. J Bacteriol. 2006, 188 (12): 4560-4569. 10.1128/JB.01908-05.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Thompson DK, Beliaev AS, Giometti CS, Tollaksen SL, Khare T, Lies DP, Nealson KH, Lim H, Yates J, Brandt CC, Tiedje JM, Zhou J: Transcriptional and proteomic analysis of a ferric uptake regulator (Fur) mutant of Shewanella oneidensis: Possible involvement of Fur in energy metabolism, transcriptional regulation, and oxidative stress. Appl Environ Microbiol. 2002, 68 (2): 881-892. 10.1128/AEM.68.2.881-892.2002.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank James K. Fredrickson for comments on the manuscript and invaluable discussions. This research was supported by The U.S. Department of Energy under the Genomics: GTL Program through Shewanella Federation, Office of Biological and Environmental Research, Office of Science. Oak Ridge National Laboratory is managed by University of Tennessee-Battelle LLC for the Department of Energy under contract DOE-AC05-00OR22725.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jizhong Zhou.

Additional information

Authors' contributions

HG carried out mutant construction and characterization, microarray analysis, weight matrix construction and screening, and drafted the manuscript. XW performed protein expression and purification, microarray analysis, and drafted the manuscript. ZKY participated in mutant construction and characterization, and microarray analysis. TP oversaw the protein work. JZ supervised all parts of the study. All authors read and approved the final manuscript.

Haichun Gao, Xiaohu Wang contributed equally to this work.

Electronic supplementary material

12864_2007_1236_MOESM1_ESM.ppt

Additional file 1: Comparison of expression measurements by microarray and Real Time qRT-PCR assays. The data provided represent the validation of microarray data by qRT-PCR assays. (PPT 124 KB)

12864_2007_1236_MOESM2_ESM.doc

Additional file 2: Genes that exhibit significant changes in the ΔarcA strain. The data provided represent all genes whose expression was altered by arcA mutation. (DOC 1 MB)

12864_2007_1236_MOESM3_ESM.doc

Additional file 3: Genes that exhibit significant changes in the ΔarcA strain (for discussion). The data provided represent all genes listed in additional file 2 that were discussed in details. (DOC 614 KB)

12864_2007_1236_MOESM4_ESM.doc

Additional file 4: Genes in operons whose upstream regions contain an ArcA binding site. The data provided represent all operons whose upstream region contains a putative ArcA binding site. (DOC 618 KB)

Additional file 5: Primers used in this study. The data provided represent all primers used in this study. (DOC 56 KB)

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Gao, H., Wang, X., Yang, Z.K. et al. Probing regulon of ArcA in Shewanella oneidensis MR-1 by integrated genomic analyses. BMC Genomics 9, 42 (2008). https://doi.org/10.1186/1471-2164-9-42

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2164-9-42

Keywords