Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2000 Dec;182(24):7075–7077. doi: 10.1128/jb.182.24.7075-7077.2000

Transcription Activation by a Variety of AraC/XylS Family Activators Does Not Depend on the Class II-Specific Activation Determinant in the N-Terminal Domain of the RNA Polymerase Alpha Subunit

Susan M Egan 1,*, Andrew J Pease 2,, Jeffrey Lang 2, Xiang Li 2, Vydehi Rao 1, William K Gillette 3,, Raquel Ruiz 1,4, Juan L Ramos 4, Richard E Wolf Jr 4
PMCID: PMC94837  PMID: 11092872

Abstract

The N-terminal domain of the RNA polymerase α subunit (α-NTD) was tested for a role in transcription activation by a variety of AraC/XylS family members. Based on substitutions at residues 162 to 165 and an extensive genetic screen we conclude that α-NTD is not an activation target for these activators.


The AraC/XylS family is a large family of transcription regulators, many of whose members activate virulence factors in bacterial pathogens and hence are of interest as potential targets of antibacterial agents (9). Virtually all AraC/XylS family members are capable of transcription activation, and thus it is likely the mechanisms used by these proteins to activate transcription have been conserved, although subsets of family members may use different mechanisms. A variety of AraC/XylS family members have been shown to require the RNA polymerase α subunit C-terminal domain (α-CTD) and the C-terminal end of the ς subunit for full activation (4, 1219; S. M. Egan and C. C. Holcroft, unpublished results; R. Ruiz, J. L. Ramos, and S. M. Egan, unpublished results). However, for several family members it is believed that one or more additional activation targets have yet to be identified. For example, SoxS has been shown to be capable of activating in vitro transcription of a class II promoter when the reconstituted RNA polymerase lacks both the α-CTD and the C-terminal residues of ς70 (K.-W. Jair and R. E. Wolf, Jr., unpublished results). There is also evidence that additional activation targets may exist in the cases of MarA, RhaS and RhaR (4, 12, 13) (Holcroft and Egan, unpublished results; R. Martin, personal communication).

Effect of α-NTD derivatives on activation by RhaS, RhaR, XylS, MarA and SoxS.

Niu et al. (24) have demonstrated that residues 162 to 165 of the RNA polymerase α subunit N-terminal domain (α-NTD) are required for transcription activation at class II cyclic AMP (cAMP) receptor protein (CRP)-dependent promoters where the CRP binding site overlaps the promoter −35 hexamer. To test whether α-NTD plays a role in transcription activation by a variety of AraC/XylS family activators, we transformed strains carrying a wild-type chromosomal copy of rpoA with plasmids overexpressing either wild-type α or previously described alanine substitution derivatives of α (24).

Activation of the rhaBAD promoter requires RhaS bound to a site that overlaps the −35 hexamer, CRP bound at −92.5, and α-CTD (6, 7, 12). Activation of the divergent rhaSR operon requires RhaR bound to a site that overlaps the −35 hexamer, CRP bound at −111.5 and α-CTD (29, 30) (Holcroft and Egan, unpublished). We grew strains carrying the α-NTD derivatives and rhaB-lacZ or rhaS-lacZ fusions in MOPS (morpholinepropanesulfonic acid) minimal medium with 0.4% glycerol, 0.2% l-rhamnose, and 125 μg of ampicillin per ml and then assayed for β-galactosidase expression as previously described (3). We found that none of the substitutions produced a significant defect in activation (Table 1).

TABLE 1.

RhaS and RhaR activation with alanine substitutions in α-NTD

α-NTD derivative β-Galactosidase sp act (% of wild-type activity)a
Φ(rhaB-lacZ)Δ110 Φ(rhaS-lacZ)Δ128
Wild type 617 ± 45 (100) 132 ± 6 (100)
162A 576 ± 51 (93) 138 ± 10 (104)
163A 566 ± 38 (92) 131 ± 10 (99)
164A 597 ± 50 (97) 135 ± 6 (102)
165A 592 ± 48 (96) 122 ± 9 (93)
162–165A 575 ± 56 (93) 113 ± 9 (86)
a

β-Galactosidase specific activity (in Miller units) was measured from single-copy rhaB-lacZ or rhaS-lacZ fusions in a wild-type strain background transformed with plasmids encoding either wild-type or substitution derivatives within the N-terminal domain of α. Φ(rhaB-lacZ)Δ110 (in strain SME1035) includes the binding sites for both the RhaS and CRP activators (7), while Φ(rhaS-lacZ)Δ128 (in strain SME2503) includes sites for both RhaR and CRP (Holcroft and Egan, unpublished). Cultures were grown in the presence of the inducer l-rhamnose. Values are the averages of three independent experiments. 

In the presence of an effector such as 3-methylbenzoate, XylS activates expression of the Pseudomonas putida TOL plasmid meta promoter, Pm, from a site that overlaps the −35 hexamer (10). The Pm promoter system was reconstituted in Escherichia coli MC4100 (28) by transformation with pERD100, which carries Φ(Pm-′lacZ) (1), a derivative of pLOW2 (11) encoding xylS, and the plasmid encoding wild-type α or alanine substitution derivatives. These strains were grown in Luria-Bertani (LB) medium with 100 μg of ampicillin, 25 μg of kanamycin, and 10 μg of tetracycline per ml, and β-galactosidase assays were performed as previously described (23, 26). We found that expression of the α-NTD derivatives had no significant effects on activation by XylS (Table 2) or the related activator (25) XylS1 (data not shown).

TABLE 2.

XylS activation at Φ(Pm-′lacZ) with alanine substitutions in α-NTD

α-NTD derivative β-Galactosidase sp act (% of wild-type activity)a
−3MBz +3MBz
Wild type 70 ± 3 (100) 1520 ± 46 (100)
162A 70 ± 2 (100) 1625 ± 42 (107)
163A 70 ± 2 (100) 1790 ± 66 (118)
164A 70 ± 2 (100) 1900 ± 15 (125)
165A 75 ± 2 (107) 1745 ± 90 (115)
162–165A 75 ± 1 (107) 1760 ± 17 (116)
a

β-Galactosidase specific activity (in Miller units) was measured from plasmid borne (Pm-′lacZ) in E. coli MC4100 transformed with pLOW2 encoding wild-type XylS and plasmids encoding either wild-type or substitution derivatives within the N-terminal domain of α. Cultures were grown either in the absence (−) or the presence (+) of the inducer 3-methylbenzoate (3MBz). Values are the averages of three independent experiments. 

The structure of the single domain MarA protein has been determined in complex with DNA (27). MarA is capable of activating transcription of a large variety of promoters (2), in some cases from a site that overlaps the −35 hexamer (class II), and in other cases from a site further upstream (class I) (20). We tested the effect of the α-NTD derivatives on MarA-dependent activation at lacZ fusions to two class I (fpr and zwf, data not shown) and three class II (inaA, fumC, and micF) promoters (Table 3). Cultures were grown in LB medium-ampicillin (100 μg/ml) and induced with 5 mM salicylate for 1 h, and β-galactosidase activity was assayed as described previously (22, 23). We found no significant defects at any of the MarA-dependent promoters.

TABLE 3.

MarA activation at class II promoters with alanine substitutions in α-NTD

α-NTD derivative β-Galactosidase sp act (% of wild-type activity)a
Φ(inaA-lacZ) Φ(fumC-lacZ) Φ(micF-lacZ)
Wild type 88/107 (100) 326/376 (100) 702/559 (100)
162A 87/110 (101) 338/309 (92) 648/556 (95)
163A 88/107 (100) 339/373 (101) 719/504 (97)
164A 95/113 (107) 303/394 (99) 686/581 (100)
165A 126/115 (124) 317/394 (101) 659/586 (99)
162–165A 102/113 (110) 312/436 (107) 669/529 (95)
a

β-Galactosidase specific activity (in Miller units) was measured from single- copy inaA-, fumC- and micF-lacZ fusions in strains N8457, N9638 and N9639 (21), respectively, transformed with plasmids encoding either wild-type or substitution derivatives within the N-terminal domain of α. Cultures were grown in the presence of the inducer salicylate. Data from experiment 1 are shown before the slash, and data from experiment 2 are shown after the slash. 

Similar to MarA, SoxS consists of a single domain and can activate class I and class II promoters (8). Activation of class II promoters was not significantly decreased upon deletion of α-CTD (15), and residues at the C-terminal end of ς70 are not essential for transcription activation by SoxS (Jair and Wolf, unpublished). To test for a role of α-NTD in SoxS activation, we assayed strains bearing translational fusions of four class II SoxS-dependent promoters (fumC, micF, nfo, and sodA) (Table 4) and two class I SoxS-dependent promoters (zwf and fpr) (data not shown). Cultures were grown in LB medium-ampicillin (125 μg/ml), induced for 1 h with 0.5 mM paraquat, and then assayed as previously described (23). The α-NTD derivatives conferred no significant effects on transcription activation of the class I or class II SoxS-dependent promoters.

TABLE 4.

SoxS activation at class II promoters ith alanine substitutions in α-NTD

Promoter rpoA mutation Activity (Miller units)a
Induction ratio % of wild-type activity
Uninduced Induced
fumC Wild type 66 2210 33 100
162A 72 2540 35 115
163A 74 2120 29 96
164A 71 2000 28 90
165A 74 2400 32 109
162–165A 82 2920 36 132
micF Wild type 43 325 7.5 100
162A 44 390 8.8 120
163A 45 335 7.4 104
164A 45 400 8.9 123
165A 47 290 6.2 89
162–165A 63 345 5.5 107
nfo Wild type 225 2400 10.6 100
162A 205 2000 9.7 84
163A 225 2070 9.2 86
164A 215 2100 9.7 88
165A 205 2460 12.1 103
162–165A 205 2440 12.0 102
sodA Wild type 1,080 9,530 8.8 100
162A 915 8,990 9.8 94
163A 1,105 8,770 7.9 92
164A 905 9,560 10.6 100
165A 880 7,930 9.0 83
162–165A 805 8,960 11.2 94
a

β-Galactosidase specific activity was measured from single-copy lacZ fusions in a wild-type strain background transformed with plasmids encoding either wild-type or substitution derivatives within the N-terminal domain of α. Cultures were grown either in the absence or the presence of the inducer paraquat. The percent wild-type value was calculated from the induced Miller unit value in the presence of the α-NTD mutant compared with the induced Miller unit value obtained with wild-type α. Values are the averages of three independent experiments. 

Genetic screen for mutations in rpoA resulting in SoxS activation defects.

Given that the mutations at residues 162 to 165 had no effect, a screen for other rpoA mutations affecting activation of class II SoxS-dependent promoters was designed. To construct the screening strain, we moved a soxR constitutive mutation (31), which provided an intermediate level of SoxS, into a strain that contained a fumC-lacZ fusion (21). This strain was transformed with derivatives of plasmid pREIIα (5) in which the entire rpoA gene had been subjected to PCR mutagenesis with Taq polymerase (32) using primers with the same sequence as those used by Niu et al. (24). The transformants were screened on lactose-tetrazolium plates (23) containing ampicillin (100 μg/ml) and kanamycin (20 μg/ml). The strain carrying the soxRC1 allele produced white colonies with light pink centers whereas a similarly uninduced isogenic strain with the wild-type allele of soxR produced red colonies. In the presence of paraquat, both strains produced white colonies. We demonstrated that less than a twofold reduction in lacZ expression in this strain resulted in reddish colonies that were clearly distinguishable from the pink-centered wild-type colonies. Approximately 24,000 transformants were screened from 26 independent mutagenesis reactions. No mutations in α-NTD were isolated that conferred a defective phenotype; however, the screen readily yielded mutations in α-CTD.

We next confined a genetic screen to α-NTD (by PCR mutagenesis of only the XbaI-to-HindIII fragment of pREIIα) to determine whether any α-NTD substitutions conferred activation defects. Twenty independent PCR mutagenesis mixtures were ligated into pREIIα, and 18,000 transformants were screened. Only one transformant with an activation-deficient phenotype was identified, and this plasmid turned out to have a rearrangement that produced an α-CTD deletion. As a control, we also screened for mutations in α-CTD and found 16 apparent activation-deficient mutants among just two independent PCRs and 2,000 transformants. Therefore, while we readily obtained mutations in the α-CTD by using this mutagenesis strategy, we were again unable to isolate any mutations in the α-NTD that influenced SoxS activation.

Remarks.

From the results of this work, we conclude that transcription activation by RhaS, RhaR, XylS, MarA, and SoxS does not require contact with the 162-to-165 determinant of α-NTD, nor, most likely, any other portion of α-NTD. The activators tested in this study represent a diverse set of AraC/XylS family proteins, which, with the exception of particularly related pairs (MarA/SoxS and RhaS/RhaR), share only 24 to 28% amino acid sequence identity. It is likely, therefore, that our conclusions apply to many other AraC/XylS family members.

Acknowledgments

We are very grateful to Richard H. Ebright for providing the plasmids encoding α-NTD derivatives and Robert Martin and Judah Rosner for providing strains.

Work in the laboratory of S.M.E. was supported by Public Health Service grant GM55099 from the National Institute of General Medical Sciences and the Franklin Murphy Molecular Biology Endowment. Work in the laboratory of R.E.W. was supported by Public Health Service grant GM27113 from the National Institutes of General Medical Sciences. R.R. received a travel fund from the Spanish Ministry of Education to visit the laboratory of S.M.E. at the University of Kansas. Work in the laboratory of J.L.R. was funded by grant BIO-97-0641.

REFERENCES

  • 1.Abril M A, Michan C, Timmis K N, Ramos J L. Regulator and enzyme specificity of TOL plasmid-encoded upper pathway for degradation of aromatic hydrocarbons and expansion of the substrate range of the pathway. J Bacteriol. 1989;171:6782–6790. doi: 10.1128/jb.171.12.6782-6790.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Barbosa T M, Levy S B. Differential expression of over 60 chromosomal genes in Escherichia coli by constitutive expression of MarA. J Bacteriol. 2000;182:3467–3474. doi: 10.1128/jb.182.12.3467-3474.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Bhende P M, Egan S M. Amino acid-DNA contacts by RhaS: an AraC family transcription activator. J Bacteriol. 1999;181:5185–5192. doi: 10.1128/jb.181.17.5185-5192.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Bhende P M, Egan S M. Genetic evidence that transcription activation by RhaS involves specific amino acid contacts with sigma 70. J Bacteriol. 2000;182:4959–4969. doi: 10.1128/jb.182.17.4959-4969.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Blatter E E, Ross W, Tang H, Gourse R L, Ebright R H. Domain organization of RNA polymerase α subunit: C-terminal 85 amino acids constitute a domain capable of dimerization and DNA binding. Cell. 1994;78:889–896. doi: 10.1016/s0092-8674(94)90682-3. [DOI] [PubMed] [Google Scholar]
  • 6.Egan S M, Schleif R F. DNA-dependent renaturation of an insoluble DNA binding protein. Identification of the RhaS binding site at rhaBAD. J Mol Biol. 1994;243:821–829. doi: 10.1006/jmbi.1994.1684. [DOI] [PubMed] [Google Scholar]
  • 7.Egan S M, Schleif R F. A regulatory cascade in the induction of rhaBAD. J Mol Biol. 1993;234:87–98. doi: 10.1006/jmbi.1993.1565. [DOI] [PubMed] [Google Scholar]
  • 8.Fawcett W P, Wolf R E., Jr Purification of a MalE-SoxS fusion protein and identification of the control sites of Escherichia coli superoxide-inducible genes. Mol Microbiol. 1994;14:669–679. doi: 10.1111/j.1365-2958.1994.tb01305.x. [DOI] [PubMed] [Google Scholar]
  • 9.Gallegos M-T, Schleif R, Bairoch A, Hofmann K, Ramos J L. AraC/XylS family of transcriptional regulators. Microbiol Mol Biol Rev. 1997;61:393–410. doi: 10.1128/mmbr.61.4.393-410.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Gonzalez-Perez M M, Ramos J L, Gallegos M T, Marques S. Critical nucleotides in the upstream region of the XylS-dependent TOL meta-cleavage pathway operon promoter as deduced from analysis of mutants. J Biol Chem. 1999;274:2286–2290. doi: 10.1074/jbc.274.4.2286. [DOI] [PubMed] [Google Scholar]
  • 11.Hansen L H, Sorensen S J, Jensen L B. Chromosomal insertion of the entire Escherichia coli lactose operon, into two strains of Pseudomonas, using a modified mini-Tn5 delivery system. Gene. 1997;186:167–173. doi: 10.1016/s0378-1119(96)00688-9. [DOI] [PubMed] [Google Scholar]
  • 12.Holcroft C C, Egan S M. Roles of cyclic AMP receptor protein and the carboxyl-terminal domain of the α subunit in transcription activation of the Escherichia coli rhaBAD operon. J Bacteriol. 2000;182:3529–3535. doi: 10.1128/jb.182.12.3529-3535.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Jair K, Martin R G, Rosner J L, Fujita N, Ishihama A, Wolf R E., Jr Purification and regulatory properties of MarA protein, a transcriptional activator of Escherichia coli multiple antibiotic and superoxide resistance promoters. J Bacteriol. 1995;177:7100–7104. doi: 10.1128/jb.177.24.7100-7104.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Jair K-M, Yu X, Skarstad K, Thony B, Fujita N, Ishihama A, Wolf R E., Jr Transcriptional activation of promoters of the superoxide and multiple antibiotic resistance regulons by Rob, a binding protein of the Escherichia coli origin of chromosomal replication. J Bacteriol. 1996;178:2507–2513. doi: 10.1128/jb.178.9.2507-2513.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Jair K-W, Fawcett W P, Fujita N, Ishihama A, Wolf R E., Jr Ambidextrous transcriptional activation by SoxS: requirement for the C-terminal domain of the RNA polymerase alpha subunit in a subset of Escherichia coli superoxide-inducible genes. Mol Microbiol. 1996;19:307–317. doi: 10.1046/j.1365-2958.1996.368893.x. [DOI] [PubMed] [Google Scholar]
  • 16.Landini P, Bown J A, Volkert M R, Busby S J W. Ada protein-RNA polymerase ς subunit interaction and α subunit-promoter DNA interactions are necessary at different steps in transcription activation at the Escherichia coli ada and aidB promoters. J Biol Chem. 1998;273:13307–13312. doi: 10.1074/jbc.273.21.13307. [DOI] [PubMed] [Google Scholar]
  • 17.Landini P, Busby S J. The Escherichia coli Ada protein can interact with two distinct determinants in the ς70 subunit of RNA polymerase according to promoter architecture: identification of the target of Ada activation at the alkA promoter. J Bacteriol. 1999;181:1524–1529. doi: 10.1128/jb.181.5.1524-1529.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Landini P, Gaal T, Ross W, Volkert M R. The RNA polymerase α subunit carboxyl-terminal domain is required for both basal and activated transcription from the alkA promoter. J Biol Chem. 1997;272:15914–15919. doi: 10.1074/jbc.272.25.15914. [DOI] [PubMed] [Google Scholar]
  • 19.Lonetto M A, Rhodius V, Lamberg K, Kiley P, Busby S, Gross C. Identification of a contact site for different transcription activators in region 4 of the Escherichia coli RNA polymerase ς70 subunit. J Mol Biol. 1998;284:1353–1365. doi: 10.1006/jmbi.1998.2268. [DOI] [PubMed] [Google Scholar]
  • 20.Martin R G, Gillette W K, Rhee S, Rosner J L. Structural requirements for marbox function in transcriptional activation of mar/sox/rob regulon promoters in Escherichia coli: sequence, orientation and spacial relationship to the core promoter. Mol Microbiol. 1999;34:431–441. doi: 10.1046/j.1365-2958.1999.01599.x. [DOI] [PubMed] [Google Scholar]
  • 21.Martin R G, Gillette W K, Rosner J L. Promoter discrimination by the related transcriptional activators MarA and SoxS: differential regulation by differential binding. Mol Microbiol. 2000;35:623–634. doi: 10.1046/j.1365-2958.2000.01732.x. [DOI] [PubMed] [Google Scholar]
  • 22.Martin R G, Rosner J L. Fis, an accessorial factor for transcriptional activation of the mar (multiple antibiotic resistance) promoter of Escherichia coli in the presence of the activator MarA, SoxS, or Rob. J Bacteriol. 1997;179:7410–7419. doi: 10.1128/jb.179.23.7410-7419.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Miller J H. Experiments in molecular genetics. Cold Spring Harbor, N.Y: Cold Spring Harbor Laboratory Press; 1972. [Google Scholar]
  • 24.Niu W, Kim Y, Tau G, Heyduk T, Ebright R H. Transcription activation at class II CAP-dependent promoters: two interactions between CAP and RNA polymerase. Cell. 1996;87:1123–1134. doi: 10.1016/s0092-8674(00)81806-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Osborne D J, Pickup R W, Williams P A. The presence of two homologous meta pathway operons on TOL plasmid pWW53. J Gen Microbiol. 1988;134:2965–2975. doi: 10.1099/00221287-134-11-2965. [DOI] [PubMed] [Google Scholar]
  • 26.Ramos J L, Stolz A, Reineke W, Timmis K N. Altered effector specificities in regulators of gene expression: TOL plasmid xylS mutants and their use to engineer expansion of the range of aromatics degraded by bacteria. Proc Natl Acad Sci USA. 1986;83:8467–8471. doi: 10.1073/pnas.83.22.8467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Rhee S, Martin R G, Rosner J L, Davies D R. A novel DNA-binding motif in MarA: the first structure for an AraC family transcriptional activator. Proc Natl Acad Sci USA. 1998;95:10413–10418. doi: 10.1073/pnas.95.18.10413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Silhavy T J, Berman M L, Enquist L W. Experiments with gene fusions. Cold Spring Harbor, N.Y: Cold Spring Harbor Laboratory Press; 1984. [Google Scholar]
  • 29.Tobin J F, Schleif R F. Purification and properties of RhaR, the positive regulator of the l-rhamnose operons of Escherichia coli. J Mol Biol. 1990;211:75–89. doi: 10.1016/0022-2836(90)90012-B. [DOI] [PubMed] [Google Scholar]
  • 30.Tobin J F, Schleif R F. Transcription from the rha operon psr promoter. J Mol Biol. 1990;211:1–4. doi: 10.1016/0022-2836(90)90003-5. [DOI] [PubMed] [Google Scholar]
  • 31.Tsaneva I R, Weiss B. soxR, a locus governing a superoxide response regulon in Escherichia coli K-12. J Bacteriol. 1990;172:4197–4205. doi: 10.1128/jb.172.8.4197-4205.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Zhou Y H, Zhang X P, Ebright R H. Random mutagenesis of gene-sized DNA molecules by use of PCR with Taq DNA polymerase. Nucleic Acids Res. 1991;19:6052. doi: 10.1093/nar/19.21.6052. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Bacteriology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES