Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Crosstalk of Escherichia coli FadR with Global Regulators in Expression of Fatty Acid Transport Genes

  • Youjun Feng,

    Affiliation Department of Microbiology, University of Illinois, Urbana, Illinois, United States of America

  • John E. Cronan

    j-cronan@life.uiuc.edu

    Affiliations Department of Microbiology, University of Illinois, Urbana, Illinois, United States of America, Department of Biochemistry, University of Illinois, Urbana, Illinois, United States of America

Abstract

Escherichia coli FadR plays two regulatory roles in fatty acid metabolism. FadR represses the fatty acid degradation (fad) system and activates the unsaturated fatty acid synthetic pathway. Cross-talk between E. coli FadR and the ArcA-ArcB oxygen-responsive two-component system was observed that resulted in diverse regulation of certain fad regulon β-oxidation genes. We have extended such analyses to the fadL and fadD genes, the protein products of which are required for long chain fatty acid transport and have also studied the role of a third global regulator, the CRP-cAMP complex. The promoters of both the fadL and fadD genes contain two experimentally validated FadR-binding sites plus binding sites for ArcA and CRP-cAMP. Despite the presence of dual binding sites FadR only modestly regulates expression of these genes, indicating that the number of binding sites does not determine regulatory strength. We report complementary in vitro and in vivo studies indicating that the CRP-cAMP complex directly activates expression of fadL and fadD as well as the β-oxidation gene, fadH. The physiological relevance of the fadL and fadD transcription data was validated by direct assays of long chain fatty acid transport.

Introduction

Much of our current knowledge of bacterial fatty acid metabolism comes from studies with Escherichia coli [1]. The fatty acid degradation (fad) pathway is primarily responsible for the transport, activation and β-oxidation of fatty acids [2]. The known long-chain fatty acid (LCFA) transport system components are the outer-membrane transport protein, FadL [3], [4] and the inner-membrane associated acyl-CoA synthetase, FadD [5], [6]. In this system, the FadL transporter delivers exogenous long chain fatty acids across the cell membrane into the periplasmic space [3], [4]. From there by an unknown mechanism the LCFA enter the cytosol where they become activated to their acyl-CoA thioesters by FadD which allows entry of the acyl chains into the β-oxidation cycle [5], [6]. E. coli FadR coordinates the catabolic and anabolic fatty acid pathways (Fig. 1). In this dual role FadR acts as a repressor for the entire set of fad regulon genes and also functions as an activator for unsaturated fatty acid biosynthesis pathway by increasing transcription of both fabA [7], [8] and fabB [9] (Fig. 1). The physiological ligands that antagonize FadR binding to its cognate operators are LCFA CoA thioesters synthesized by the cytosolic acyl-CoA synthetase, FadD [7], [10]. In strains lacking either FadD or FadL addition of LCFA fails to induce fad regulon expression because the regulatory ligand cannot be synthesized. Therefore, extremely stringent regulation of fadD and fadL genes is precluded because this would block derepression.

thumbnail
Figure 1. Regulation of fatty acid metabolism by E. coli FadR.

A. In the absence of a long chain acyl-CoA E. coli FadR represses the fad regulon genes [8], [23] whereas it activates transcription of fabA and fabB, the two genes of the unsaturated fatty acid synthetic pathway [7][9]. B. Binding of long chain acyl-CoA species results in large changes in FadR structure resulting in dissociation of the protein from its operator sites. Dissociation increases fad regulon expression whereas expression of fabA and fabB is decreased. The ovals denote FadR whereas the small open circles represent the acyl-CoA pool.

https://doi.org/10.1371/journal.pone.0046275.g001

Similar to most of the other fad regulon genes, fadL [4], [11], [12] and fadD [12], [13] are controlled at the transcriptional level by two different regulatory systems in addition to FadR, the oxygen-sensitive ArcA-ArcB two-component system and the cyclic AMP (cAMP) receptor protein-cyclic AMP (CRP-cAMP) complex. Our interest in fadL and fadD arose from the presence of two FadR-binding sites in the promoters of both genes whereas the other FadR regulon fad genes each have only a single site. This raised the question of regulatory interactions among the multiple regulators, FadR, ArcA and CRP-cAMP when bound to the respective promoter regions. Cho and coworkers [12] initially used quantitative RT-PCR to assay regulation of fadL and fadD by FadR and ArcA under anaerobic conditions and reported that deletion of either arcA, fadR alone, or both arcA and fadR resulted in increased fadL expression by 34-, 4- or 69-fold, respectively, whereas fadD transcription increased 69-, 4.5-, and 84-fold. These workers did not investigate the effects of CRP-cAMP. Classical catabolite repression of the fad pathway has long been known. Pauli and coworkers [14] reported that Fad enzyme levels were very low in wild type cells grown in glucose even in the presence of inducing levels of fatty acid (or in constitutive fadR mutant strains) and that glucose inhibition could be partially relieved by addition of cAMP. CRP mutant strains also had low fad enzyme activities. In this laboratory Clark [15] using early lacZ fusion technology showed that glucose acted at the transcriptional level in fadBA and fadE expression. Although as discussed below CRP-cAMP binding sites for several fad genes have been proposed, none had been directly tested for the ability to bind the complex. Therefore, it remained possible that CRP-cAMP regulation of the fad regulon was an indirect effect.

We report experiments defining the interactions of these regulatory proteins and their effects on fatty acid transport.

Materials and Methods

Bacterial strains and growth conditions

All the strains used here are E. coli K-12 derivatives (Table 1). The bacterial media were used for bacterial growth and analyses of β-galactosidase activity. These included LB medium (Luria-Bertani medium containing 10 g of tryptone, 5 g of yeast extract and 10 g of NaCl per liter), Rich broth (RB medium containing 10 g of tryptone, 1 g of yeast extract, and 5 g of NaCl per liter) and minimal medium M9 [16] supplemented with 0.4% glucose or other carbon source, 0.1% Vitamin-Free Casamino Acids, 0.1 mM CaCl2, and 0.001% thiamine. When necessary, antibiotics were used as follows (in mg/liter): sodium ampicillin, 100; kanamycin sulfate, 50; chloramphenicol, 20 and tetracycline HCl, 15. To monitor transcriptional regulation by ArcA-ArcB, an anaerobic environment (5% H2, 75% N2, and 20% CO2) was generated in an anaerobic environmental system (Bio-Bag environmental chamber type A; Becton Dickinson). The E. coli strains were grown on modified LB media in which potassium nitrate (5 mM) was added as an electron acceptor [17]. Fatty acids such as oleic acid (Sigma) were neutralized with KOH, solubilized with Tergitol NP-40, and used as an inducer at 5 mM final concentration.

Plasmids and genetic manipulations

The pCR2.1-TOPO vector (Invitrogen) was used for PCR cloning and sequencing whereas expression vector pET28a (Novagen) was used for protein preparation (Table 2). All plasmids constructed were validated by PCR analyses plus direct DNA sequencing.

Both FYJ159 and FYJ295 are, respectively, derivatives of E. coli strain MC4100, in which either a fadD–lacZ transcriptional fusion [18] or a fadL-lacZ transcriptional fusion, was integrated into the chromosomal attλ site (Table 1). Strain FYJ104 which carries a chromosomal fadL-lacZ transcription fusion (Table 1) was constructed using the FLP-mediated site-specific recombination method [19]. The kanamycin resistance cassette was removed from the fadL::km strain JW2341-1 (Table 1) by expression of the FLP recombinase encoded by plasmid pCP20 (Table 1) to give strain FYJ103, which retained a single FLP recombinase target (FRT) site. The FRT site was used for site-specific integration of the lacZ fusion plasmid, pKG37 (an improved version of pCE71) containing a FRT site upstream of a promoterless lacZY genes, a kanamycin resistance gene, and the R6K origin of replication (Table 1) [19]. The transformants were screened on LB agar plates containing kanamycin and 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside (X-Gal) at 37°C to obtain the chromosomal fadL-lacZ fusion strain FYJ104. The fusion plasmid was stably integrated due to its R6K origin and loss of the temperature-sensitive pCP20 plasmid [19]. The fadL promoter-lacZ junction was validated by a PCR using primers, fadL-P plus lacZ-R (Table 2).

P1vir phage transduction

P1vir transductions were carried out as described by Miller [20] with minor modifications. Strains FYJ133, FYJ134, FYJ141, FYJ142, FYJ169 and FYJ170 were generated by transduction of strains FYJ59 (ΔarcA), FYJ60-2 (Δcrp), FYJ139 (ΔarcA ΔfadR::Tn10), FYJ140 (Δcrp ΔfadR::Tn10), FYJ77 (Δcrp ΔarcA), and FYJ78 (Δcrp ΔarcA ΔfadR::Tn10), respectively (Table 1) with a P1vir lysate grown on FYJ104 (fadL-lacZ) with selection for kanamycin resistance. Similarly, a P1vir lysate grown on FYJ159 (fadD-lacZ) was used for transduction of the following strains FYJ57, FYJ59, FYJ60-2, FYJ139, FYJ140, FYJ77 and FYJ80 with selection for kanamycin resistance to give strains FYJ161, FYJ162, FYJ163, FYJ164, FYJ165, FYJ166 and FYJ167, respectively (Table 1). Transduction of strains FYJ104, FYJ59 and FYJ60-2 with a P1vir lysate grown on MFH8 (fadR::Tn10) with selection for tetracycline resistance gave strains FYJ105 (ΔfadR::Tn10, fadL-lacZ), FYJ139 (ΔarcA ΔfadR::Tn10) and FYJ140 (Δcrp ΔfadR::Tn10), respectively.

β-Galactosidase assays

Mid-log phase cultures in LB, RB or minimal media (with or without supplementation with various carbon sources), were collected by centrifugation, washed twice with Z Buffer [21] and assayed for β-galactosidase activity after lysis with sodium dodecyl sulfate-chloroform [21]. The data were recorded in triplicate with no less than three independent experiments.

Protein expression and purification

Hexahistidine-tagged E. coli FadR (and/or CRP) proteins were produced in E. coli BL21 (DE3) carrying the expression plasmid pET28-fadRec (and/or pET28-crp) (Table 1) by induction of bacterial cultures at an OD600 nm of 0.8–1.0 with 0.3 mM IPTG at 30°C for 3 h [17], [22]. The cells were pelleted washed twice with ice cold PBS buffer (101.4 mM Na2HPO4, 1.8 mM KH2PO4, 137 mM NaCl, 2.7 mM KCl, 8% glycerol, pH 7.4), dissolved in the same buffer and lysed using a French pressure cell. The extracts were centrifuged to remove bacterial debris and the supernatants loaded onto a nickel chelate column (Qiagen). Following washing with ten column volumes of with PBS buffer containing 50 mM imidazole, the FadR proteins were eluted with 150 mM imidazole. Appropriate eluted protein fractions were pooled and dialyzed against PBS buffer then concentrated by ultrafiltration (30 kDa cut-off, Amicon Ultra) [17]. The protein purity was judged by 12% SDS-PAGE, followed by staining with Coomassie brilliant blue R250 (Sigma, St. Louis, MO). Both FadR and CRP proteins were characterized by liquid chromatography quadrupole time-of-flight mass spectrometry of tryptic peptides and chemical cross-linking as described previously [18].

Electrophoretic mobility shift assays

These assays of the interaction between the fadD and fadL promoters, FadR and the cAMP-CRP complex were done essentially as previously reported [18], [23]. All of the FadR (and/or CRP) probes were prepared by annealing two complementary primers (Table 2) by incubation in TEN buffer (10 mM Tris-HCl, 1 mM EDTA, 100 mM NaCl, pH 8.0) at 95°C for 5 min followed by slow cooling to 25°C and then DIG labeling by terminal transferase with DIG-ddUTP (Roche). DNA probes (Table 2) for assay of DNA binding by the CRP-cAMP complex were similarly synthesized. The digoxigenin-labeled DNA probes (either 0.1 or 0.2 pmol) were incubated with either DNA binding protein in binding buffer (Roche) for 15 min at room temperature and then analyzed by native PAGE (6.5% PAGE for the CRP probes and 7% PAGE for all other probes). The separations were then visualized as previously described [18], [23].

Fatty acid transport assays

Fatty acid transport was assayed as described by Klein et al. [24] with minor modifications. To avoid complications by β-oxidation strains that carried a fadBA disruption (strains FYJ82, FYJ83 and SI203) were used (Table 1). To test ArcA-P regulation of fatty acid transport, FYJ82 strain (ΔarcA) was compared with the wild type strain SI203 (Table 1). Overnight cultures were inoculated into 10 ml of RB liquid media supplemented with potassium nitrate (5 mM) as electron acceptor and kept in a fully anaerobic chamber at 37°C for ∼10 hrs [17]. The anaerobic environment (5% H2, 75% N2, and 20% CO2) was generated by an anaerobic environmental chamber (Bio-Bag type A, Becton Dickinson) [17]. Cultures in exponential phase were treated with 100 mM chloramphenicol for 10 min prior to assay. 1-14C-Oleic acid (American Radio-labeled Chemicals) was injected into the anaerobic bacterial cultures using a syringe fitted with a fine needle (30G1 PrecisionGlide) to a final concentration of 45.5 µM. The cultures were mixed well by vigorous vortex mixing and incubated at room temperature for about 15 min with anaerobiosis monitored by an anaerobic indicator (0.001% resazurin). Finally, the bacteria were collected by centrifugation (4.200× g, 16 min), and washed for five times with iced RB medium. One ml of the cultures were subjected to membrane phospholipid extraction [18], [25] after measuring culture absorbance (A600) and adjusted to an absorbance of 1.5. The phospholipids were then separated from any residual fatty acid by thin layer chromatography [18], [25]. Assay of the effects of the CRP-cAMP system on logarithmic phase cultures aerobically grown in LB liquid proceeded in a similar manner.

Bioinformatic analyses

The known or predicted DNA binding sites recognized by either FadR or ArcA (or CRP) were all from the E. coli literature. Multiple alignments were done using ClustalW2 (http://www.ebi.ac.uk/Tools/clustalw2/index.html), and the resultant output was processed by program ESPript 2.2 (http://espript.ibcp.fr/ESPript/cgi-bin/ESPript.cgi), generating the final BLAST version.

Results and Discussion

Regulatory complexity in the LCFA transport promoters, fadL and fadD

The fadL and fadD genes encode the proteins known to be required for LCFA transport (Fig. 2A). The fadL and fadD transcriptional start sites are located 95 bp and 60 bp upstream of the translation start sites, respectively [6], [26] (Fig. 2B). Despite a seemingly straightforward role in metabolism, both promoters contain demonstrated or annotated binding sites for three different transcription factors, FadR, ArcA and CRP (http://www.ecocyc.org). Moreover, fadL and fadD promoters are the only fad regulon promoters that contain two distinct FadR binding sites [6], [26]. In the fadL promoter the FadR sites are separated by only 8 bp whereas in the fadD promoter the two FadR sites are separated by 68 bp (Fig. 2B and C). The locations of the ArcA binding sites determined by Cho and coworkers [12] also differ. The fadL ArcA site is 5 bp from FadR site 1 whereas the fadD ArcA site overlaps FadR site 2 by one bp, raising the possibility of crosstalk between the two repressor proteins (Fig. 2). The fadL CRP site was reported to lie downstream of the transcription start site [4] (Fig. 2B), a position incompatible with the usual activator function of CRP. Searches using the CRP consensus sequence of Zheng et al. [27] produced another fadL candidate binding site upstream of the transcription start site. Although this seemed a much more plausible position for activation of transcription, experimental verification of the CRP binding site was required. Indeed, although CRP-cAMP regulation of the fad regulon genes at the physiological level has been known for many years [14], none of the proposed sites have been experimentally validated.

thumbnail
Figure 2. The fadL and fadD promoters of E. coli.

A. Genomic contexts of fadL and fadD on the E. coli chromosome (the two genes are separated by 570 kb). The transcription start sites (S) are indicated by arrows. Dotted lines denote spanning with long distance. B and C. Sequences of the fadL promoter and the fadD promoter, respectively. All the demonstrated sites bound by either FadR or ArcA are labeled with black rectangles, whereas the annotated CRP binding sites that were validated in this study are labeled with blue rectangles. A dotted rectangle labels a fadL CRP binding site incorrectly predicted by EcoCyc (http://www.ecocyc.org). White letters shaded in red denote the identical residues, red letters shaded in white mean similar bases, whereas white-shaded black letters denote unrelated. Dots denote gaps. “M” denotes the translation initiation site. The numbers are given relative to the transcription start site (S). The fadL and fadD FadR binding sites and their relative spacing are largely conserved in the genomes of Citrobacter, Kelbsiella and Enterobacter.

https://doi.org/10.1371/journal.pone.0046275.g002

The two FadR-binding sites of fadL and fadD result in only modest repression

To obtain a parallel comparison of the FadR binding sites of the two promoters we used electrophoretic mobiltity shift analyses (EMSA) to assay the binding abilities of the individual sites over a range of FadR concentrations (Fig. 3). The two fadL FadR sites bound FadR with equivalent affinities (essentially complete binding of the probes by 50 nM FadR, Fig. 3B and D). In contrast, although fadD site 1 showed a binding affinity comparable to that of the fadL sites (Fig. 3F), FadR binding by fadD site 2 (Fig. 3H) was >10-fold weaker than the other three sites (Fig. 3B, D and F). Addition of 25–50 µM oleoyl-CoA resulted in loss of FadR binding by all four sites (Fig. 3C, E, G and I). Therefore, our data are in good agreement with the reported DNase I foot-printing results [6], [11] and provide data on the effects of the acyl-CoA regulatory ligand that is lacking in the foot-printing experiments. To monitor expression of fadL and fadD, each promoter was fused to a LacZ reporter gene to allow expression to be assayed by β-galactosidase activity (Fig. 4). Deletion of FadR resulted in only modest derepression of the two LCFA transport system genes (2 to 2.5 fold for fadL and 2 to 3-fold for fadD) in medium with acetate as sole carbon source (Fig. 4A and B) whereas expression of fadBA and fadH increased by 5 to10-fold (Fig. 4C and D). Increased expression of genes fadL and fadD in the ΔfadR strain was also seen when the carbon source was either glucose or glycerol (Fig. 4A and B), These observations are generally consistent with those obtained upon oleate induction of strains carrying a functional FadR (Fig. 4E).

thumbnail
Figure 3. E. coli fadL and fadD both carry two functional FadR-binding sites.

A. sequence alignments of several known FadR binding sites from E. coli fad regulon. White letters with red background denote strictly conserved bases whereas yellow background letters denote highly conserved residues. The dual FadR sites of fadL and fadD are highlighted in bold italics. B and D. Gel shift assays of FadR binding to both fadL promoter sites. F and H. Gel shift assays of FadR binding to both fadD promoter sites. Long chain fatty acyl-CoA species block binding of FadR to the two fadL sites (C and E), as well as to the two fadD sites (G and I). FadR was used at concentrations of 0 (denoted by a minus sign), 1, 2, 5, or 10 pmol. FadR was incubated with 0.1 pmol of DIG-labelled probe in a total volume of 10 µl. For the acyl-CoA experiments the components were: probe, 0.1 pmol; FadR, 1 pmol and acyl-CoA, 50 pmol. Designations: C9:0; nonanoyl–CoA; C10:0, decanoyl-CoA; C16:0, palmitoyl-CoA; C16:1, palmitoleoyl-CoA; C18:0, stearoyl-CoA; C18:1, oleoyl-CoA.

https://doi.org/10.1371/journal.pone.0046275.g003

thumbnail
Figure 4. FadR repression and oleate induction of fatty acid transport system.

A and B. Repression of fadL and fadD by FadR in cultures grown on various carbon sources. C and D. Parallel experiments with two β-oxidation genes fadBA and fadH are given for comparison. Strains FYJ104 (wild type) FYJ105 (ΔfadR::Tn10) were used for fadL whereas strains FYJ159 (wild type) and FYJ161 (ΔfadR) were used for fadD. Strains SI203 (wild type) and SI207 (ΔfadR) were used for fadBA whereas strains FYJ55 (wild type) and FYJ56 (ΔfadR) were used for fadH. E. Induction of fadL and fadD expression by oleate. Oleate induction of the fatty acid transport system genes was compared with those of both fadBA and fadH expression. Oleic acid (5 mM) was added as the sole carbon source and compared to 5 mM acetate. Four strains expressing wild type FadR (FYJ295 (fadL-lacZ), FYJ159 (fadD-lacZ), SI203 (fadBA-lacZ), and FYJ55 (fadH-lacZ), were used. All strains were grown under aerobic condition (10 ml of culture in a 50 ml flask shaken at 200 rpm at 37°C). ß -Galactosidase (ß-gal) assays were conducted in triplicate and the error bars indicate standard deviations.

https://doi.org/10.1371/journal.pone.0046275.g004

Expression of fadL and fadD is directly activated by the CRP-cAMP complex

The global regulator cAMP-CRP complex [28] can act as either an activator [29] or a repressor [30] in expression of genes involved in many E. coli metabolic pathways. Although putative class I cAMP-CRP binding sites had been proposed upstream of some fad regulon genes and activation of fadH transcription was observed in vivo [17], direct physical evidence for DNA binding of these promoters by the cAMP-CRP complex was lacking. We therefore used EMSAs with purified CRP in the presence or absence of cAMP to test the proposed sites. SDS-PAGE analyses indicated that the apo-form of CRP of molecular weight ∼23 kDa was homogenous (Fig. 5A). Chemical cross-linking showed the protein was dimeric as previously reported [28] (Fig. 5B). Liquid chromatography mass spectrometry analysis of tryptic peptides demonstrated that our recombinant protein matched E. coli CRP with 66% coverage (Fig. 5C).

thumbnail
Figure 5. Expression and characterization of E. coli CRP.

A. SDS-PAGE profile of purified CRP. B. Chemical cross-linking assays of purified E. coli CRP. The minus sign decodes no addition of the chemical cross-linker EGS, whereas the EGS concentrations were 1, 2, 5, 10, 15, and 20 µM. The cross-linking reaction mixtures were loaded on 12% SDS-PAGE. M denotes the Pre-stained broad range protein standards (BioRad). C. MS identification of E. coli CRP protein. The tryptic peptides matching the CRP sequence are given in bold underlined type.

https://doi.org/10.1371/journal.pone.0046275.g005

Higashitani and coworkers [4] predicted a fadL CRP binding site (Fig. 2B) centered at 57 bp downstream of the fadL transcription start site (termed fadL2) that is covered by our fadL probe 2, Fig. 6C). In contrast, we favored a site centered 157 bp upstream (Fig. 2C and 6A) covered by fadL probe 1 (Fig. 6B). In the fadH promoter, we previously proposed a CRP binding site (called fadH site 1 covered by fadH probe 1, Fig. 6A and E) and subsequently have predicted a second site (called fadH2 and covered by fadH probe 2, Fig. 6A and F). Given that sequence alignments of these sites showed only six conserved bp (Fig. 6A), the function of these sites required direct testing by EMSAs. As expected from numerous prior investigations (e.g., Lawson et al. [31]), CRP lacked DNA binding activity in the absence of its cAMP ligand (Fig. 6). Gel shift assays confirmed that the CRP-cAMP complex efficiently bound fadL probe 1 (Fig. 6B), but failed to bind fadL probe 2 (Fig. 6C). Therefore, fadL site 1 is a functional CRP-cAMP complex binding site whereas the site predicted by Higashitani and coworkers [4] is nonfunctional. Unlike Class II sites which overlap the −35 hexamer and Class III sites which have tandem CRP binding sites, the position of fadL site 1 identifies it as an atypical site that is located too far upstream to be reached the C-terminal domain (CTD) of the α–subunit of RNA polymerase which seems to have maximum stretch of 90–120 bp upstream of the promoter. In contrast the demonstrated fadD CRP-cAMP complex binding site [27] can be clearly identified as a Class I site (Fig. 2C and 6D). Both proposed fadH promoter sites bound the CRP-cAMP complex and therefore comprise a bona fide Class III site of the type that is composed of two Class I sites (some Class III sites have a Class II site coupled to a Class I site) (Fig. 6E and F).

thumbnail
Figure 6. The cAMP-CRP complex binds the fadL, fadD and fadH promoters.

A. Sequence analyses of the CRP binding sites of fadL, fadD and fadH. White letters shaded in black denote strictly conserved bases, bold letters in grey represent highly similar residues, and dots mean gaps. The newly deduced CRP-binding site (fadL1 of A) binds the cAMP-CRP complex (panel B) whereas the CRP-binding site predicted earlier [17] (abbreviated fadL2) does not (panel C). D Binding of the fadD promoter region by the cAMP-CRP complex. E and F. The fadH promoter region contains two functional CRP-binding sites, fadH1 and fadH2, respectively. The DIG-labeled probe shifted by the cAMP-CRP complex is indicated by an arrow. All EMSA experiments were carried out using 6.5% native PAGE and representative results are shown. The protein samples were incubated with 0.6 pmol of DIG-labeled probe in a total volume of 20 µl that contained 200 pmol cAMP (when added). The right hand four lanes of each contained (left to right) 10, 5, 10 and 20 pmol of CRP, respectively.

https://doi.org/10.1371/journal.pone.0046275.g006

To evaluate the effects of CRP on expression of the fatty acid transport genes, in vivo assays of β-galactosidase activities of the fadD and fadL lacZ fusions in Δcrp mutant strains were compared to strains expressing wild type CRP (Fig. 7). As a control we also assayed fadH expression and found that expression decreased in the absence of CRP (Fig. 6E and F) (Fig. 7D). As expected, the activities of both transport promoters were modestly decreased upon the loss of crp (Fig. 7A and B) verifying that CRP activates fadL and fadD transcription. Although in agreement with others (http://www.ecocyc.org), we failed to identify any CRP-binding site candidates in the fadBA promoter, expression of fadBA in the crp deletion strain is decreased relative to the wild type strain (Fig. 7C). This may indicate an indirect activation of the fadBA promoter by the CRP-cAMP complex. It seems possible that the indirect activation is due to action of the FIS protein, expression of which is activated by CRP binding to three discontinuous sites [27]. Indeed, a putative FIS binding site (TTGCATATTTTTAACACAA, −36 to 18) lies within the fadBA promoter and FIS is a known activator of fadBA transcription [32].

thumbnail
Figure 7. Activation of fatty acid transport system gene expression by the CRP-cAMP complex.

A and B. Decreased expression levels of fadL and fadD, respectively, were seen in the absence of CRP-cAMP. Strains FYJ104 (wild type) and FYJ134 (Δcrp) carry the fadL-lacZ transcriptional fusion whereas FYJ159 (wild type) and FYJ65 (Δcrp) contain the fadD-lacZ transcriptional fusion. C and D. Expression of fadBA and fadH are positively regulated by the functional CRP-cAMP complex. Strains SI203 (wild type) and FYJ83 (Δcrp) carry the fadBA-lacZ transcriptional fusion whereas FYJ55 (wild type) and FYJ65 (Δcrp) contain the fadH-lacZ transcriptional fusion. The cultures were grown in either LB or RB media. beta-Galactosidase activities were from at least three independent experiments, and the error bars indicate standard deviations.

https://doi.org/10.1371/journal.pone.0046275.g007

Expression of fadL and fadD are additively repressed by FadR and ArcA under anaerobic conditions

The weak effects of FadR inactivation on expression of fadL and fadD suggested that other regulatory proteins may be involved and thus we tested the effects of deletion of the genes encoding the other proteins proposed to bind these promoters. The oxygen-sensitive two-component system ArcA-ArcB was reported to negatively regulate transcription of several fad regulon genes, including fadL, fadD, fadBA and fadH [12]. That report [12] also contained in vitro and in vivo evidence that the phosphorylated ArcA regulator (ArcA-P) bound the promoters of the target genes (Fig. 8A). We recently reported cross-talk between FadR and ArcA in expression of other fad regulon genes and found that the interaction could be either additive (fadH) (Fig. 8E) or synergistic (fadBA) (Fig. 8D) [17]. The fadD promoter FadR site 2 and the ArcA site overlap by a single bp (Fig. 2C) as also seen in fadH [17] whereas in the fadL promoter the ArcA and FadR site 1 sequences are separated by 5 bp (Fig. 2B). Thus, cross-talk between FadR and ArcA in fadL and/or fadD expression seemed likely.

thumbnail
Figure 8. Negative regulation of fadL and fadD by ArcA-P under anaerobic conditions.

A. Alignments of the ArcA-P binding sites of with those of other fad regulon genes. Red background indicates strictly conserved bases whereas yellow background denotes less conserved bases. B. Modulation of LacZ activity of the fadBA-lacZ transcriptional fusion by different regulatory proteins under anaerobic conditions. The fadBA-lacZ strains were SI203 (WT), SI207 (ΔfadR), FYJ82 (ΔarcA), FYJ83 (Δcrp), FYJ84 (ΔarcA ΔfadR), FYJ85 (Δcrp ΔfadR), FYJ86 (Δcrp ΔarcA) and FYJ87 (Δcrp ΔarcA ΔfadR), respectively. C. Transcriptional control of the fadH-lacZ fusion by different regulatory proteins under anaerobic conditions. The fadH-lacZ fusion strains were FYJ55 (WT), FYJ56 (ΔfadR), FYJ64 (ΔarcA), FYJ65 (Δcrp), FYJ68 (ΔarcA ΔfadR), FYJ81 (Δcrp ΔfadR), FYJ79 (Δcrp ΔarcA) and FYJ80 (Δcrp ΔarcA ΔfadR), respectively. D. Effect on fadL transcriptional levels by different regulatory proteins under anaerobic conditions. The fadL-lacZ were FYJ104 (WT), FYJ105 (ΔfadR), FYJ133 (ΔarcA), FYJ134 (Δcrp), FYJ141 (ΔarcA ΔfadR), FYJ142 (Δcrp ΔfadR), FYJ169 (Δcrp ΔarcA), and FYJ170 (Δcrp ΔarcA ΔfadR), respectively. E. Regulation of fadD transcription by different regulatory proteins under anaerobic conditions. The fadD-lacZ fusion strains were FYJ159 (WT), FYJ161 (ΔfadR), FYJ162 (ΔarcA), FYJ163 (Δcrp), FYJ164 (ΔarcA ΔfadR), FYJ165 (Δcrp ΔfadR), FYJ166 (Δcrp ΔarcA), and FYJ167 (Δcrp ΔarcA ΔfadR), respectively. The strains were grown on LB agar plates supplemented with 5 mM nitrate (KNO3) as the electron receptor. Anaerobic environments were generated using Bio-Bag environmental chamber type A as described [17]. beta-Galactosidase activities were recorded from at least six independent assays and are expressed as means ± standard deviations. *, P<0.005; **, P<0.001.

https://doi.org/10.1371/journal.pone.0046275.g008

To test for cross-talk strains carrying arcA and/or fadR null mutations plus the chromosomal fadL-lacZ (or fadD-lacZ) transcriptional fusions were constructed. In general agreement with the report of Cho et al. [12], we found that the absence of arcA under anaerobic conditions resulted in increased transcription of fadL and fadD by 2 to 3-fold and >20-fold, respectively, (Fig. 8B and C). Our levels of derepression are somewhat less than those reported by Cho and coworkers [12] which we attribute to the gene used by these workers as their internal reference [17], [33]. However, our data differed from those of Cho and coworkers in the effects reported on FadR regulation of fadL transcription under anaerobic conditions when ArcA is functional (we see no effect versus their reported 4-fold increase). Although our statistical analysis argued that these two regulatory proteins control transcription of fadL and fadD in an additive manner, the wide variations in the data and the overlapping error bars provide caveats to this interpretation.

In the absence of the cAMP receptor protein (CRP) complex (cAMP-CRP), FadR and ArcA-ArcB repression of fadL and fadD expression was relatively weak (Fig. 8) and similar to that seen with fadBA and fadH [17]. Thus it seems that regulation by both FadR and ArcA-ArcB rely on the cAMP-CRP complex to activate transcription such that it can be further modulated.

Physiological relevance of ArcA and the CRP-cAMP complex to fatty acid transport

Uptake of [1-14C]oleic acid was assayed to evaluate the physiological consequences of ArcA-ArcB and the CRP-cAMP complex on fatty acid transport (Fig. 9). To rule out potential interference by β-oxidation, the strains studied lacked fadBA (Table 1). Incorporation of [1-14C]oleic acid into the membrane phospholipids of the ΔarcA strain (FYJ82) was consistently >2-fold greater that that of its parental strain under anaerobic conditions (Fig. 9A). Further analyses by thin layer chromatography showed that incorporation into the three major membrane phospholipids phosphatidylethanolamine (PE), phosphatidylglycerol (PG) and cardiolipin (CL) (Fig. 9B), consistent with anaerobic repression of fadL (and/or fadD) by ArcA-P (Fig. 8B and C).

thumbnail
Figure 9. Effects of ArcA and CRP on incorporation of [1-14C]oleic acid into membrane phospholipid.

A. Incorporation in the ΔarcA strain, FYJ82, and the wild type strain, SI203. At least three independent experiments were carried out and the data are expressed as mean ± standard deviation. B. A representative autoradiogram of a TLC separation of the 14C-labeled phospholipids of a panel A experiment. The phospholipid species are phosphatidylethanolamine (PE), phosphatidylglycerol (PG) and cardiolipin (CL). A representative autoradiogram is given. C. Incorporation in the Δcrp strain, FYJ83 and the wild type strain, SI203. At least five independent experiments were performed and the data are expressed in means ± SD. D. A representative autoradiogram of a TLC separation of the 14C-labeled phospholipids of a panel C experiment. Bacterial strains used in panel A and B were kept under anaerobic condition (details seen in Materials and Methods), whereas experiments in panel C and D were routinely conducted under aerobic condition.

https://doi.org/10.1371/journal.pone.0046275.g009

In agreement with the observed CRP-cAMP activation of fadL and fadD transcription (Fig. 7A and B), quantitative determination of total 14C-labeled bacterial membrane phospholipids showed that the level of oleic acid incorporation in the strain lacking crp is only about one third that of the parental strain (Fig. 9C and D). To our knowledge these are the first direct physiological data that directly show that these two global regulators modulate LCFA transport pathway in E. coli.

Conclusions

Transcriptional regulation of fatty acid transport in E. coli involves three distinct regulatory systems, the specialized FadR system and the two global systems, ArcA and CRP-cAMP. FadR action is straightforward, it acts as a classical LacI-type repressor and only weakly represses fadD and fadL expression. CRP-cAMP regulation is also straightforward, fatty acids are a low status carbon source and E. coli prefers to use the highest status carbon source, glucose. In the presence of glucose (or in the absence of CRP), the other two regulators have little transcription to regulate. The apparent surprise is the stringent repression by ArcA seen under anaerobic conditions even in the absence of FadR because E. coli has a pathway to degrade fatty acids under anaerobic conditions [34] and low expression of FadD and FadL would seem likely to compromise function of the anaerobic pathway. However, under anaerobic conditions a new acyl-CoA synthetase, FadK, is induced that replaces FadD [35]. Unlike FadD which is inactive with short chain fatty acids [24], [35], FadK strongly prefers short chain length acids and such acids are the preferred growth substrates of the anaerobic β-oxidation pathway [34] although LCFA can also be utilized. Short chain acids readily enter fadL strains [36] and thus decreased expression of FadL is irrelevant for these growth substrates. However, degradation of LCFA such as oleate via the anaerobic pathway should require FadL. These apparent contradictions can be reconciled by our finding that ArcA represses fadL expression less than fadD expression (only about 4-fold) (Fig. 8B and C) and that FadK activates oleate poorly [35]. Hence, the significant level of FadL expressed under ArcA repression probably provides sufficient oleate transport to satisfy the poor catalytic activity of FadK with this substrate and allow the observed slow anaerobic growth on oleate [35].

Author Contributions

Conceived and designed the experiments: YF JC. Performed the experiments: YF. Analyzed the data: YF JC. Contributed reagents/materials/analysis tools: YF JC. Wrote the paper: YF JC.

References

  1. 1. Clark D, Cronan J (2005) Two-carbon compounds and fatty acids as carbon sources. In A Bock, R. Curtiss III, J Karper, P. Karp, F. Neidhardt, T. Nystrom, J. Slauch, C. Squiress, and D. Ussery (ed.), EcoSal–Escherichia coli and Salmonella: cellular and molecular biology. http://www.ecosal.org. ASM Press, Washington, DC.
  2. 2. Black PN, DiRusso CC (1994) Molecular and biochemical analyses of fatty acid transport, metabolism, and gene regulation in Escherichia coli. Biochim Biophys Acta 1210: 123–145.
  3. 3. van den Berg B, Black PN, Clemons WM Jr, Rapoport TA (2004) Crystal structure of the long-chain fatty acid transporter FadL. Science 304: 1506–1509.
  4. 4. Higashitani A, Nishimura Y, Hara H, Aiba H, Mizuno T, et al. (1993) Osmoregulation of the fatty acid receptor gene fadL in Escherichia coli. Mol Gen Genet 240: 339–347.
  5. 5. Fulda M, Heinz E, Wolter FP (1994) The fadD gene of Escherichia coli K12 is located close to rnd at 39.6 min of the chromosomal map and is a new member of the AMP-binding protein family. Mol Gen Genet 242: 241–249.
  6. 6. Black PN, DiRusso CC, Metzger AK, Heimert TL (1992) Cloning, sequencing, and expression of the fadD gene of Escherichia coli encoding acyl coenzyme A synthetase. J Biol Chem 267: 25513–25520.
  7. 7. Henry MF, Cronan JE Jr (1992) A new mechanism of transcriptional regulation: release of an activator triggered by small molecule binding. Cell 70: 671–679.
  8. 8. Henry MF, Cronan JE Jr (1991) Escherichia coli transcription factor that both activates fatty acid synthesis and represses fatty acid degradation. J Mol Biol 222: 843–849.
  9. 9. Campbell JW, Cronan JE Jr (2001) Escherichia coli FadR positively regulates transcription of the fabB fatty acid biosynthetic gene. J Bacteriol 183: 5982–5990.
  10. 10. Cronan JE Jr (1997) In vivo evidence that acyl coenzyme A regulates DNA binding by the Escherichia coli FadR global transcription factor. J Bacteriol 179: 1819–1823.
  11. 11. DiRusso CC, Metzger AK, Heimert TL (1993) Regulation of transcription of genes required for fatty acid transport and unsaturated fatty acid biosynthesis in Escherichia coli by FadR. Mol Microbiol 7: 311–322.
  12. 12. Cho BK, Knight EM, Palsson BO (2006) Transcriptional regulation of the fad regulon genes of Escherichia coli by ArcA. Microbiology 152: 2207–2219.
  13. 13. Farewell A, Diez AA, DiRusso CC, Nystrom T (1996) Role of the Escherichia coli FadR regulator in stasis survival and growth phase-dependent expression of the uspA, fad, and fab genes. J Bacteriol 178: 6443–6450.
  14. 14. Pauli G, Ehring R, Overath P (1974) Fatty acid degradation in Escherichia coli: requirement of cyclic adenosine monophosphate and cyclic adenosine monophosphate receptor protein for enzyme synthesis. J Bacteriol 117: 1178–1183.
  15. 15. Clark D (1981) Regulation of fatty acid degradation in Escherichia coli: analysis by operon fusion. J Bacteriol 148: 521–526.
  16. 16. Iram SH, Cronan JE (2006) The β-oxidation systems of Escherichia coli and Salmonella enterica are not functionally equivalent. J Bacteriol 188: 599–608.
  17. 17. Feng Y, Cronan JE (2010) Overlapping repressor binding sites result in additive regulation of Escherichia coli FadH by FadR and ArcA. J Bacteriol 192: 4289–4299.
  18. 18. Feng Y, Cronan JE (2011) Complex binding of the FabR repressor of bacterial unsaturated fatty acid biosynthesis to its cognate promoters. Mol Microbiol 80: 195–218.
  19. 19. Ellermeier CD, Janakiraman A, Slauch JM (2002) Construction of targeted single copy lac fusions using lambda Red and FLP-mediated site-specific recombination in bacteria. Gene 290: 153–161.
  20. 20. Miller JH (1992) A short course in bacterial genetics. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y. .
  21. 21. Miller J (1972) Experiments in Molecular Genetics. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.
  22. 22. Feng Y, Cronan JE (2011) The Vibrio cholerae fatty acid regulatory protein, FadR, represses transcription of plsB, the gene encoding the first enzyme of membrane phospholipid biosynthesis. Mol Microbiol 81: 1020–1033.
  23. 23. Feng Y, Cronan JE (2009) A new member of the Escherichia coli fad regulon: transcriptional regulation of fadM (ybaW). J Bacteriol 191: 6320–6328.
  24. 24. Klein K, Steinberg R, Fiethen B, Overath P (1971) Fatty acid degradation in Escherichia coli. An inducible system for the uptake of fatty acids and further characterization of old mutants. Eur J Biochem 19: 442–450.
  25. 25. Feng Y, Cronan JE (2009) Escherichia coli unsaturated fatty acid synthesis: complex transcription of the fabA gene and in vivo identification of the essential reaction catalyzed by FabB. J Biol Chem 284: 29526–29535.
  26. 26. Black PN (1991) Primary sequence of the Escherichia coli fadL gene encoding an outer membrane protein required for long-chain fatty acid transport. J Bacteriol 173: 435–442.
  27. 27. Zheng D, Constantinidou C, Hobman JL, Minchin SD (2004) Identification of the CRP regulon using in vitro and in vivo transcriptional profiling. Nucleic Acids Res 32: 5874–5893.
  28. 28. Harman JG (2001) Allosteric regulation of the cAMP receptor protein. Biochim Biophys Acta 1547: 1–17.
  29. 29. van Aalten DM, DiRusso CC, Knudsen J (2001) The structural basis of acyl coenzyme A-dependent regulation of the transcription factor FadR. Embo J 20: 2041–2050.
  30. 30. Husnain SI, Busby SJ, Thomas MS (2009) Downregulation of the Escherichia coli guaB promoter by upstream-bound cyclic AMP receptor protein. J Bacteriol 191: 6094–6104.
  31. 31. Lawson CL, Swigon D, Murakami KS, Darst SA, Berman HM, et al. (2004) Catabolite activator protein: DNA binding and transcription activation. Curr Opin Struct Biol 14: 10–20.
  32. 32. Bradley MD, Beach MB, de Koning AP, Pratt TS, Osuna R (2007) Effects of Fis on Escherichia coli gene expression during different growth stages. Microbiology 153: 2922–2940.
  33. 33. Zhang Y, Cronan JE Jr (1996) Polar allele duplication for transcriptional analysis of consecutive essential genes: application to a cluster of Escherichia coli fatty acid biosynthetic genes. J Bacteriol 178: 3614–3620.
  34. 34. Campbell JW, Morgan-Kiss RM, Cronan JE Jr (2003) A new Escherichia coli metabolic competency: growth on fatty acids by a novel anaerobic β-oxidation pathway. Mol Microbiol 47: 793–805.
  35. 35. Morgan-Kiss RM, Cronan JE (2004) The Escherichia coli fadK (ydiD) gene encodes an anerobically regulated short chain acyl-CoA synthetase. J Biol Chem 279: 37324–37333.
  36. 36. Nunn WD, Giffin K, Clark D, Cronan JE Jr (1983) Role for fadR in unsaturated fatty acid biosynthesis in Escherichia coli. J Bacteriol 154: 554–560.
  37. 37. Haldimann A, Wanner BL (2001) Conditional-replication, integration, excision, and retrieval plasmid-host systems for gene structure-function studies of bacteria. J Bacteriol 183: 6384–6393.
  38. 38. Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, et al. (2006) Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol Syst Biol 2: 2006 0008.
  39. 39. Iram SH, Cronan JE (2005) Unexpected functional diversity among FadR fatty acid transcriptional regulatory proteins. J Biol Chem 280: 32148–32156.
  40. 40. 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.
  41. 41. Cherepanov PP, Wackernagel W (1995) Gene disruption in Escherichia coli: TcR and KmR cassettes with the option of Flp-catalyzed excision of the antibiotic-resistance determinant. Gene 158: 9–14.