Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jul 3.
Published in final edited form as: Nat Rev Microbiol. 2008 Jun 3;6(7):507–519. doi: 10.1038/nrmicro1912

Advances in bacterial promoter recognition and its control by factors that do not bind DNA

Shanil P Haugen 1, Wilma Ross 1, Richard L Gourse 1
PMCID: PMC3700611  NIHMSID: NIHMS478987  PMID: 18521075

Abstract

Early work identified two promoter regions, the −10 and −35 elements, that interact sequence specifically with bacterial RNA polymerase (RNAP). However, we now know that several additional promoter elements contact RNAP and influence transcription initiation. Furthermore, our picture of promoter control has evolved beyond one in which regulation results solely from activators and repressors that bind to DNA sequences near the RNAP binding site: many important transcription factors bind directly to RNAP without binding to DNA. These factors can target promoters by affecting specific kinetic steps on the pathway to open complex formation, thereby regulating RNA output from specific promoters.


Because many, and perhaps most, gene products are conserved among different species of bacteria, the extraordinary variety of bacterial life must result from differences in the relative amounts of these products and in the timing of their expression. Regulation can occur at every step on the pathway to gene expression, but transcription initiation is probably the most frequently regulated step. Differences in intrinsic promoter strength and in the action of transcription factors that regulate promoters are sufficient to account for much of the variation in the levels of RNA that are produced by different operons13.

In contrast to archaea and eukaryotes, bacteria contain only one form of RNA polymerase (RNAP) core enzyme (E), which is a complex of five subunits (α2ββ′ω). However, in most bacterial species, there are multiple forms of the σ-factor specificity subunit, and thus multiple forms of RNAP holoenzyme (referred to as Eσ). Some differences in promoter use and activity result from changes in the availability of different RNAP holoenzymes under different nutritional or environmental conditions. However, the strengths of different promoters that are recognized by the same form of holoenzyme, for example Eσ70 (the main holoenzyme in Escherichia coli), can vary over many orders of magnitude, with some promoters producing less than one RNA copy per cell generation and other promoters producing tens of thousands of RNA copies per cell generation13.

Alignments of E. coli promoter sequences from as long ago as 1975 identified the −10 and −35 hexamers, the primary determinants of basal promoter strength1,2. Furthermore, several additional promoter elements that affect strength and regulation have been identified since that time46 (FIG. 1). Sequence variation within individual promoter elements affects specific interactions of the promoter with two RNAP subunits, σ and α, and results in the vast range of basal promoter strengths that are found in nature14. At many or most promoters, basal promoter strength is increased or decreased by activator or repressor proteins that bind to specific DNA sequences near to, or overlapping, the RNAP binding site, thereby restricting the effects of these regulators to only those promoters that contain the transcription-factor binding sites3.

Figure 1. DNA elements and RNAP modules that contribute to promoter recognition by Eσ70.

Figure 1

The optimal UP element (α binding site) is an alternating series of A and T tracts (−57 to −38; 5′-AAAWWTWTTTTNNNAAANNN-3′; W = A or T; and N = any base)37. −35 refers to the −35 element, consensus sequence 5′-TTGACA-3′ from −35 to −30 (REF. 1). Ext refers to the extended −10 element, consensus sequence 5′-TGTG-3′ from −17 to −14 (REF. 50). −10 refers to the −10 element, consensus sequence 5′-TATAAT-3′ from −12 to −7 (REFS 1,2). Dis refers to the discriminator region, optimal sequence 5′-GGG-3′ from −6 to −4 (REFS 6,32,68). +1 is the transcription start site. The segments of the σ and α subunits of RNA polymerase (RNAP) that interact with these elements are described in the main text and REFS 1,2,6,37. αCTD, α carboxy-terminal domain.

In this Review, we first provide a brief overview of RNAP structure and the sequence of events that occur during bacterial transcription initiation. We then review recent advances in our understanding of the individual interactions between RNAP and DNA that participate in promoter recognition. Finally, we discuss the regulation of transcription initiation, in which we emphasize mechanisms other than those of the classical activators and repressors, which have been reviewed elsewhere (for a recent review, see REF. 3). The unconventional regulators described in this Review include small proteins, RNAs, nucleotides, nucleotide analogues and ions. The amounts and/or activities of these regulators change according to nutritional and environmental conditions. Similar to the conventional transcription factors, the effects of these factors are promoter specific, but they differ from conventional transcription factors in that promoter specificity is not achieved by binding to DNA7.

Transcription initiation

RNAP structure and the kinetics of the reaction

The six subunits of the holoenzyme α2ββ′ωσ′ are highly conserved in bacteria, and all except σ have homologues in archaea and eukaryotes. The α dimer forms the scaffold on which β and β′, the catalytic subunits, assemble, and the ω subunit assists β′ binding to the α2 sub-assembly1,8. The overall structure of the bacterial core enzyme9 is similar to that of the archaeal RNAP10 and the eukaryotic RNAPs11, and the only major differences are on the surface of the complex911. In both the core and holoenzyme structures of bacterial RNAP, the overall architecture resembles that of a crab claw, in which β forms one pincer and β′ the other1215 (FIG. 2a). In the holoenzyme–DNA complex, σ is oriented with the carboxy (C)-terminal domain upstream and the amino (N)-terminal domain downstream with respect to the promoter14. The conformation of σ is extended, with the domains spread primarily along one face of RNAP (BOX 1; FIG. 2b, c). The interface between σ and core RNAP is extensive, comprising more than 8,200 Å2 of surface12.

Figure 2. Models of the open complex.

Figure 2

a, b | Promoter elements and the regions of Eσ70 RNA polymerase (RNAP) that recognize promoter elements, the secondary channel and some other features of the enzyme are labelled. Segments of RNAP that bind to promoters are shown in space-fill, and ribbon representations of the DNA sections that are bound by these parts of RNAP are represented by matching colours (β, blue; β′, pink; α amino-terminal domain (αNTD), light grey; and ω, dark grey. The structures are adapted from the model of the open complex19, which is based on the crystal structure of the fork–junction complex14. The paths of the individual strands downstream from the −10 hexamer are based on modelling and biochemical studies17,19, as DNA downstream of the −10 element is not present in the fork–junction complex. c | α carboxy-terminal domains (αCTDs) and upstream DNA are included. The αCTD structure is from REF. 40. The positions of the αCTDs and the path of the DNA upstream of the −35 hexamer are based on modelling and biochemical studies19,30. Region 1.1 of σ was not resolved in the structure, but would be outside the main channel in the open complex21. NTS, non-transcribed strand; TS, transcribed strand. Coordinates for the model from REF. 19 were courtesy of R. Ebright and C. Lawson, Rutgers University, New Jersey, USA.

Box 1. σ factors.

Box 1

Most bacteria encode multiple σ factors, and, in some species, there are >60 (REF. 3). Each of the seven different σ factors inEscherichia coli, which range from 173 to 613 amino acids in length, recognizes its own cognate consensus promoter sequence. The most abundant σ factor in the bacterial cell, referred to as σ70 inE. coliA in most other species), is responsible for recognizing most of the cell’s promoters. σ70 contains four regions of conserved amino acid sequences, σ1–σ4, that are further divided into subregions1 (see the figure). In addition to region 1.1 and the nonconserved region (NCR; a segment of σ70 that, even in all members of the σ70 class, is not always present and the sequence and length of which varies widely79), these regions of sequence conservation are also present in the alternative σ factors, but not in the σ54 class (the most divergent of the σ factors1,84). In the main text, we refer primarily to theE. coli70 holoenzyme for brevity. However, the mechanisms of the alternative holoenzymes are also probably similar to that of Eσ70, with the exception of Eσ54. Core RNA polymerase (RNAP) is limiting, so the different affinities of the σ factors for core RNAP, as well as dramatic variation in the amounts of the different σ factors under different nutritional and environmental conditions, contribute to the regulation of promoter utilization by the different holoenzymes.

There are two major channels in RNAP, which are bifurcated by a long, evolutionarily conserved feature of the enzyme, the bridge helix (FIG. 2a). These channels provide solvent access to the active site in the absence of DNA. However, promoter DNA fills the main channel that is formed by the cleft between the β and β′ pincers. Therefore, it is thought that the much narrower secondary channel provides the usual route whereby nucleoside 5′-triphosphates (NTPs) access the active site16 (BOX 2).

Box 2. Factors that bind in the secondary channel of RNAP.

Box 2

The secondary channel of RNA polymerase (RNAP) is too narrow to accommodate double-stranded DNA (see the figure). However, it provides access for solvent and nucleoside 5′-triphosphates (NTPs) to the enzyme’s active site and is wide enough to accommodate back-tracked RNA (which occurs when the RNA 3′ end is no longer aligned with the active site)17. In addition, the secondary channel has emerged in recent years as a route by which small molecules and small proteins can access the RNAP active site and exert a wide range of effects ontranscription7,16,106,107.

The structural similarity between DksA and the Gre factors107,119,120 raised the possibility that Gre factors also affect ribosomal RNA (rRNA) transcription initiation. However, GreA does not mimic the effects of DksAin vitronor compensate for the absence of DksAin vivo, even when produced at concentrations that are sufficient to compensate for its lower concentration and its apparently weaker affinity for RNAP than DksA122. Thus, binding in the secondary channel of RNAP is insufficient to fully explain the effect of DksA on transcription initiation. By contrast, GreB and DksA have similar effects on rRNA promoter complexesin vitro, but the GreB concentration is much lower than that of DksA, which explains the absence of an effect ofgreBmutants on rRNA transcriptionin vivo122. When GreA was overexpressed in an otherwise wild-type strain, however, it competed with DksA and prevented DksA from modulating rRNA transcription146.

There is no DksA homologue inThermus thermophilus,but it has been proposed that another Gre factor homologue, Gfh1, acts as an anti-Gre factor by binding in the RNAP secondary channel in this bacterium147. The secondary channel is also the target of antibiotics, such as microcin J25 (REF. 148). Figure modified, with permission, from REF. 15 © (2003) Elsevier Science and REF. 149 © (2004) Elsevier Science.

The X-ray crystal structure of the promoter–Thermus aquaticus RNAP complex, known as the fork–junction complex14, provided valuable insights about the placement of DNA within RNAP. Although the 6.5 Å resolution of the fork–junction complex was insufficient to discern details of the protein–protein and protein–DNA interactions in the promoter complex, the availability of data from biochemical and genetic studies1,2,17, higher resolution structures of individual components of the promoter complex (reviewed in REFS 9,12) and a high resolution structure of the elongation complex18 make it possible to construct high-quality models of the open complex from the low-resolution fork–junction structure14,15,19.

Nevertheless, it should be emphasized that the structure is a snapshot of only one step in a multistep process and components that are not present in this structure have important roles in transcription initiation. For example, the ~85-residue C-terminal domains of the two α subunits20 and the highly acidic ~90-amino-acid N-terminal domain of σ (region 1.1)21, although modelled in some panels of FIGS 2,3 are not resolved in the structure, probably because they are mobile. The complex captured in the crystal also lacks DNA upstream of the −35 hexamer, downstream of the −10 hexamer on the non-template strand and downstream of position −12 on the template strand. Therefore, the picture of transcription initiation that was afforded by the fork–junction complex is incomplete. It should be noted that the lengths of the DNA spacers that separate different promoter elements vary from promoter to promoter, which can lead to ambiguity in nomenclature. For consistency, we use numbering that corresponds to the most frequent situation, in which the transcription start site (+1) is the seventh position downstream from the −10 hexamer, and there is a 17 bp spacer between the −10 and the −35 hexamers.

Because formation of a complex that is capable of initiating transcription is a multistep process, we need to correlate structural descriptions of complexes with specific kinetic intermediates1,2,22. It is generally assumed that there is a common mechanism that is shared by all promoters1. However, there are difficulties inherent in the detection of rapidly forming, transient complexes. Furthermore, the correlation of specific structures with specific steps in the mechanism has been complicated by variation in the precise upstream and downstream end-points of promoter complexes that have been assigned by footprinting and other biochemical assays1,2,4,2229. Some variation results from differences in the structural details of complexes that are formed by different promoters, but effects of differences in template topology (for example, DNA supercoiling), ionic conditions and temperature have also contributed to the difficulty in correlating specific structures with experimentally derived kinetically significant intermediates (for a review, see REF. 1).

For simplicity, we refer to three intermediate complexes of RNAP with promoter DNA (RP), RP closed (RPC), RP intermediate (RPI) and RP open (RPO) (FIG. 3), but we recognize that there is some ambiguity in these terms. The energetics of this process have been described as “bind, nucleate, melt” (REFS 1,2,22). The movements of the transcriptional machine that occur on the pathway to initiation are driven not by the hydrolysis of ATP, but by binding free energy — the establishment of interfaces and conformations in earlier intermediates that trigger subsequent rearrangements to form later intermediates22. Ultimately, these successive rearrangements result in DNA opening and alignment of the start-site base of the template strand with the catalytic site of the enzyme. The largest of the conformational changes in the enzyme (in terms of the burial of polar amide surface area) that occurs on the pathway to formation of the transcriptionally competent complex appears to take place after DNA-strand opening and could correspond to the clamping of the enzyme’s jaw (or jaws) onto the downstream DNA27.

Figure 3. Steps in transcription initiation.

Figure 3

In a, RPC refers to the earliest promoter complex with RNA polymerase (RNAP), although this complex is not always kinetically detectable at all promoters. RPO refers to the final complex before nucleoside 5′-triphosphate occupancy of RNAP. RPI is used as an abbreviation for all the intermediates between RPC and RPO. The structures of these intermediates remain to be defined and could even differ at different promoters. ka is the composite association-rate constant for RPO formation. kd is the composite dissociation constant. Individual steps are shown in bf (REFS 15,30), and are described in more detail in the main text. The template DNA strand is in dark green; the non-template DNA strand is in light green; and the RNA transcript that emerges from the RNA channel is in red. The catalytic Mg2+ ion in the active site is indicated by a yellow sphere. Red circles represent acidic residues in regions 1.1 and 3.2, which are mobile modules in σ70. In RPC (b), the DNA is double stranded and downstream DNA has not moved into the main channel. In RPI (c), σ region 1.1 begins to move out of, and downstream DNA begins to move into, the main channel. This probably occurs concurrent with DNA-strand separation22,27, but the intermediates (RPI) remain to be determined. In RPO (d), the DNA strands have separated, with the template strand moving into position for base pairing with the first nucleoside 5′-triphosphate (NTP). In the ‘scrunched complex’ (REFS 34,35) (e), RNA synthesis has begun, and the DNA strands in or downstream of the −10 hexamer are temporarily ‘extruded’ out of their respective channels before promoter clearance. An NTP is shown entering the enzyme through the secondary channel. In the transcription elongation complex (TEC) (f), σ has dissociated from core RNAP. Figure adapted, with permission, from REF. 15 (2003) Elsevier Science.

In the first step in the general scheme (FIG. 3a), RNAP binds to promoter DNA to form the closed complex RPC (FIG. 3b). In this complex, RNAP protects DNA in footprints from approximately −55 to approximately +1 relative to the transcription start site, and the DNA is completely double stranded1,2,2224. Although RNAP protects both the −35 hexamer and the −10 hexamer in these footprints, the DNA has not yet entered the main channel22,24,27. The complex then undergoes large changes that are referred to as isomerizations, in which the conformations of both RNAP and DNA change1,2,22,27 (discussed below). The downstream boundary of the intermediate complex RPI extends beyond the transcription start site, perhaps as far as +12 (FIG. 3c). In this complex, the DNA strands are still base paired1,2,15,22,24,30, although there could be a few transiently opened base pairs within the −10 hexamer. Ultimately, in the open complex RPO, the DNA strands are separated from approximately −11 to approximately +3 and the footprint extends downstream to approximately +20, leaving a base on the template strand (designated +1) that is able to pair with the first NTP (initiating NTP (iNTP))1,2 (FIG. 3d).

As discussed below, although DNA downstream of the −10 hexamer is not in the crystallized fork–junction complex, crosslinking experiments indicate that the non-template strand immediately downstream of the −10 hexamer is held in a groove that is formed between β and σ(REFS 17,26), where it can interact with a highly conserved module in σ region 1.2 (REFS 6,31,32) (BOX 1; FIG. 2b). The template strand threads through RNAP to the active site, and the DNA strands re-anneal at approximately +4 (REFS 2,33). Modelling based on biochemical data and on the structure of the elongation complex suggests that the double-stranded DNA downstream to approximately +12 is housed in a protein tunnel that is formed by β and β′ (REFS 14,1719,26).

The N-terminal domain of σ region 1.1 is disordered in all the X-ray structures, but biophysical approaches have shown that σ region 1.1 is in the main DNA channel of the holoenzyme (FIG. 3b) and is displaced by double-stranded DNA during open complex formation21 (FIG. 3d). The molecular details of this process remain to be determined. However, the end result of the conformational changes that contribute to melting of the −10 hexamer is an approximately 90° change in the trajectory of the DNA15,30 (FIG. 3d).

Once the open complex forms, NTP incorporation drives the transcription reaction forward. However, at most promoters, RNAP synthesizes short, abortive products before transitioning to the elongation complex. During this time, the leading edge and the active site of the enzyme move forward, but the contacts between the trailing edge of RNAP and the −35 hexamer remain intact. It appears that both DNA strands in the vicinity of the −10 hexamer are extruded from the main DNA channel onto the surface of the enzyme during this transition in a process that is called scrunching34,35 (FIG. 3e). It is proposed that the energy stored in this scrunched intermediate is used to break the interactions between RNAP and the promoter, thereby allowing RNAP to begin the transition to the elongation phase of transcription (FIG. 3f). Ultimately, formation of the elongation complex involves staged disruption of a series of contacts between σ and core RNAP33,36 (discussed below).

In the discussion that follows, we describe interactions between individual promoter DNA elements and modules of RNAP in more detail, moving from upstream to downstream with respect to the transcription start site (FIG. 1).

αCTD–DNA interactions and promoter recognition

Although the two α subunits are identical in amino acid sequence, they are not functionally equivalent, as the core enzyme is asymmetric: one α subunit interacts primarily with β and one interacts primarily with β′ (REF. 9) (FIG. 2c). Each α subunit is comprised of two domains that are tethered by a flexible linker of ~15 amino acid residues20. The α N-terminal domain (αNTD) provides the dimerization interface as well as the scaffold for core-enzyme assembly9, whereas the α C-terminal domain (αCTD) binds to DNA4. DNA sequences that result in specific binding by the αCTDs upstream of the −35 element are referred to as UP elements4,37. UP elements are widely distributed in bacterial, plasmid and phage promoters. Extensive biochemical and genetic analyses of the αCTD–DNA interaction and a high resolution X-ray structure provide a cohesive picture of the complex, in which αCTD interacts in and across the DNA minor groove using a helix–hairpin–helix motif 3740.

The two αCTDs increase promoter activity most when bound to the two DNA minor grooves just upstream of the −35 hexamer4,37 (FIGS 1,2c), but the flexible tether that links the two domains of α, coupled with bending of the DNA, can allow the αCTDs to interact with DNA that is as much as 5–6 minor grooves upstream of the −35 hexamer26. DNA binding by a single αCTD to half of an UP element (an UP element subsite) can also stimulate transcription, but to a lesser extent than binding to a full UP element37.

αCTD can also interact with DNA sequences in a base-nonspecific manner (that is, with the DNA backbone)41. The kinetic consequences of sequence-specific and nonspecific αCTD binding to DNA for transcription initiation have been investigated. Unexpectedly, not all of the effects are attributable to initial recruitment of RNAP to the promoter4143. Elimination of αCTD (or truncation of DNA upstream of the −35 hexamer) not only decreases KB, the equilibrium binding constant for the first kinetically significant complex, but it also slows down ki, the composite rate of isomerization, by at least fourfold. Exactly how αCTD affects ki remains unclear. One model is that αCTD allosterically affects a part of RNAP that contributes to the isomerization step41. An alternative model is that DNA upstream of the αCTD binding site (or sites) directly contacts a part of RNAP that is involved in the isomerization step (or steps), and that this is facilitated by the αCTD–DNA interaction28,43.

σ region 4 interactions with the −35 element and with the αCTD

Extensive genetic, biochemical and structural information has provided a detailed picture of recognition of the −35 element by a helix–turn–helix motif in σ region 4.2 (REFS 1,2,44,45) (FIG. 2a). The helix–turn–helix contacts backbone positions and bases on both the template and non-template strands, which are among the first sequence-specific interactions made between RNAP and the promoter29,46,47. When bound to the proximal UP-element subsite that is centred at approximately −41, αCTD can interact directly with σ70 region 4.2, which is bound to the −35 element48,49 (FIG. 2c). The αCTD–σ region 4.2 interface appears to differ slightly depending on whether CRP (cyclic AMP receptor protein), which activates transcription through contact with the αCTD, is bound to a DNA site that is centred at −61.5 (REFS 48,49). This suggests that transcription factors remodel the αCTD–σ region 4.2 interaction.

Extended −10 element interactions with σ region 3.0

Surface-exposed residues in σ region 3.0, an α helix that is perpendicular to region 2.4, bind in the major groove of the extended −10 element, which is separated from the −10 element by one base5,1214 (FIGS 1,2b, c). This interaction is crucial for transcription from a subset of promoters with poor matches to the −35 or −10 consensus hexamers5,5052. Although only a subset of E. coli promoters have consensus-extended −10 elements, these are more frequent in Bacillus subtilis5354. The kinetic effects of the extended −10 interaction with σ region 3.0 have not been characterized extensively, but the interaction seems to stimulate transcription, both by increasing the rate of association and by stabilizing the promoter complex6,53,54. The identity of promoter position −13, the position between the extended −10 and −10 elements, also contributes to specific recognition by at least some RNAP holoenzymes, particularly EσS (REF. 55).

The interaction between the −10 hexamer and σ region 2

The literature on the −10 hexamer–σ region 2 interaction is too detailed and extensive to be covered here in depth (see, for example, REFS 1,2). To summarize, the interactions between σ region 2, the most highly conserved region in σ, and the −10 hexamer are complex, as this element is recognized by RNAP, first, in the double-stranded DNA form and, after melting, as single-stranded DNA5660 (FIG. 2b). Sections of σ regions 2.3 and 2.4 form a continuous α helix that interacts with non-template strand bases in the −10 hexamer, and aromatic side chains in these sections of σ probably stack on bases in the −10 element during DNA-strand opening14,24,5661. The non-template adenine at −11 seems to be the most crucial base for the nucleation of strand separation. Localized melting may involve base flipping at −11, and stacking of an aromatic residue in region 2.3 on the base at −12 is proposed to prevent DNA unwinding further upstream14,6066. Unsurprisingly, because the −10 hexamer is contacted in both the double-stranded and single-stranded DNA forms, mutations in the −10 hexamer affect both KB and ki (REF. 67).

Discriminator interactions with σ region 1.2

The term discriminator was first used by Andrew Travers68 to describe a G+C-rich region between the −10 hexamer and the transcription start site in ribosomal RNA (rRNA) and transfer RNA (tRNA) promoters. Like the UP element and the extended −10 element, the base sequence in this region is not highly conserved in E. coli promoters and was therefore not originally recognized as a sequence-specific recognition element for RNAP. However, recent work has shown that σ region 1.2 contacts the non-template strand base two positions downstream from the −10 hexamer (−5 at most promoters)6,31 (FIG. 2a). Promoter sequences that contain a guanine at −5 are the most favourable for the interaction with E. coli σ70 (REFS 6,31). The guanine at −5 crosslinks to the N-terminal residues of two perpendicular α helices in E. coli σ70 region 1.2, probably with residues Y101 and/or M102 (REF. 31). The discriminator–σ region 1.2 interaction affects the rate of dissociation of RPO to RPC (REF. 6), which is a major determinant of the susceptibility of rRNA and other promoters to negative regulation by ppGpp and DksA (discussed below). Consistent with a contact between σ region 1.2 in E. coli RNAP and a guanine in the discriminator, the major σ factor in T. aquaticus binds best to a DNA sequence that contains three guanines immediately downstream of the −10 hexamer32.

Promoter escape and σ release

The σ cycle paradigm, in which σ is released upon promoter clearance and is then recycled for use by other core RNAP enzymes69, has been questioned in recent years70,71. Although σ release is not obligatory for promoter clearance72, it now seems clear that σ is usually released in a stochastic manner within the first few hundred base pairs of the start site7374 (as proposed more than 20 years ago75).

Because the interaction of σ with core RNAP to form the holoenzyme is a multistep process7678, it is not surprising that dissociation of σ from core RNAP also takes place in stages (reviewed in REF. 72). As discussed above, displacement of σ region 1.1 from the main channel of the enzyme occurs on the pathway to RPO formation to allow entry of double-stranded DNA14,15,21,30 (FIG. 3b–d). As the nascent transcript grows to ~6 nucleotides, interactions of the σ3–σ4 linker within the RNA exit channel are displaced (FIG. 3e), and σ region 4 contacts are disrupted from the β flap at a length of ~16–17 nucleotides (that is, when the transcript reaches the end of the RNA exit channel)36. The interaction of σ region 2 with the coiled-coil region of β′ is sufficient for the association of σ with core RNAP78. Furthermore, the interaction of σ region 2.2 with the coiled-coil itself could be the last contact to be disrupted to release σ from core RNAP. It also may be sufficient for the retention (or reassociation) of σ in some proportion of the elongation complexes72. The non-conserved region (σNCR) (BOX 1) that is located between σ region 1.2 and σ region 2.1 (and/or RNAP-associated transcription factors) could have roles in disruption of the σ region 2.2 interaction with the β′ coiled coil74,79.

Classical regulators of transcription

Classical repressors (for example, the Lac and λ repressors) usually prevent binding of RNAP by occluding the promoter in the vicinity of the −10 and −35 region3, but a repressor has also recently been identified that inhibits transcription solely by competing for an UP element80. Repression mechanisms have also been reported in which proteins do not prevent RNAP binding, but instead block formation of intermediates in the initiation pathway8183. Classical activators typically bind near RNAP, contacting αCTD or σ region 4.2 (REF. 3) and thereby recruiting RNAP to the promoter or facilitating a conformational change later on the pathway to initiation. Several different activators and co-activators often function together on the same promoter3.

These conventional activators and repressors share an important property: they restrict their effects to specific promoters by binding to specific DNA sequences that are near to or overlap the promoter. Below, we focus on regulators of transcription initiation that bind to RNAP and achieve specificity without binding to DNA. Some other transcription factors (reviewed in REF. 84), including many activators of Eσ54-dependent promoters, as well as the activator of Eσ70, bacteriophage N4 single-stranded DNA-binding protein, bind to DNA, but differ in their mechanisms of action from either conventional regulators or the regulators that are described below.

Regulators that do not bind DNA

Anti-σ and anti-α factors

Anti-σ factors are proteins that regulate transcription by sequestering one or more of the σ surfaces that bind to core RNAP or to promoter DNA85. Anti-σ factors are σ specific (for example, anti-σF, anti-σE or anti-σ32), and are widespread in the bacterial kingdom. The mechanisms of action of several members of this class have been defined at the atomic level86.

The phage T4-encoded regulator AsiA uses a slightly different strategy. It targets specific promoter subsets by binding to σ rather than to DNA, and it functions by interfering with −35 element recognition, both by preventing the interaction of the β flap with σ region 4 (REF. 87) and by deforming the DNA-binding surface of σ region 4 (REF. 88). In conjunction with the T4-encoded transcription factor MotA, AsiA mediates a global change in gene expression from T4 early genes to T4 middle genes.

Whereas the anti-σ factors target promoters that use a specific σ factor, the B. subtilis protein Spx modulates the function of a different RNAP subunit — α. Spx functions as an anti-α factor by sequestering activator and σ-region-4 binding surfaces on αCTD, and it can activate or inhibit transcription at specific promoters during oxidative or disulphide stress conditions89,90.

A factor that stimulates holoenzyme assembly?

Crl, a 15.8 kDa protein first identified in E. coli and found in many gammaproteobacteria, is required for efficient expression of some EσS-dependent promoters9193. Crl interacts directly with σS (and perhaps other σ factors) and increases transcription from a subset of promoters in vivo and in vitro9196. Because stimulation of transcription by Crl was much greater when it was added to σ before assembly of σ with core RNAP, and Crl stimulated transcription only when the σ concentration was low, it was proposed that Crl aids binding of σ to core RNAP93. This model is consistent with analyses of EσS formation in extracts from wild-type and crl mutant strains92. One attractive model is that Crl serves as a chaperone that helps fold or unfold σ to facilitate its binding to the core93.

As indicated above, Crl does not increase transcription in vivo from every promoter that is recognized by EσS, and it does not bind to DNA. Because Crl seems to increase the effective concentration of RNAP, it was suggested that its promoter specificity results from differences in the intrinsic binding constants of different promoters for EσS (REF. 93). Further studies are required to determine whether Crl-activated promoters are simply those that bind EσS most weakly and are therefore most sensitive to alterations in EσS concentrations.

Global negative control by ppGpp and DksA

The unusual nucleotides guanosine tetraphosphate and pentaphosphate (collectively referred to here as ppGpp) were identified approximately 40 years ago (reviewed in REFS 7,97). ppGpp is distributed widely in bacteria and is also found in chloroplasts, which evolved from bacteria97,98. Control of transcription by ppGpp has been described in many bacterial species97, and its effects on transcription have been characterized on a genome-wide scale in both E. coli99,100 and B. subtilis101. ppGpp directly and specifically inhibits promoters for rRNAs and most tRNAs7,97,102,103, as well as some promoters for mRNAs (for example, pyrBI104 and fis105). It was recently discovered that ppGpp has a cofactor in E. coli and many other bacterial species: the 151 amino acid protein DksA106, although DksA is not as widely distributed as ppGpp7,107. In addition to inhibiting some promoters, ppGpp and DksA activate a number of other promoters directly and/or indirectly7,108111 (discussed below).

The discovery that DksA mediates the effect of ppGpp on transcription in E. coli resolved the discrepancy between the 2- to 3-fold inhibition of rRNA promoters by ppGpp that has been observed in vitro in reactions that contain only RNAP and promoter DNA, and the ~20-fold inhibition that has been observed in cells102,103,106,112. Inactivation of the dksA gene virtually abolishes inhibition of rRNA promoters by ppGpp in vivo, thereby preventing the regulation of ribosome synthesis106. Elimination or overexpression of dksA has been reported to affect not only ribosome synthesis, but also several other functions, such as amino acid biosynthesis, chaperone function, cell division, quorum sensing, phage sensitivity, responses to envelope stress and virulence in a range of bacterial species (reviewed in REF. 7).

There is a strong correlation between the short lifetime of the competitor-resistant promoter complex and negative regulation by ppGpp and DksA6,7 — the competitors for RNAP binding that are used in such studies are typically either the polyanion heparin or double-stranded promoter DNA. rRNA promoter complexes are extremely short-lived, with the equilibrium between competitor-resistant complexes (for example, RPO) and competitor-sensitive intermediates (for example, RPC) becoming shifted in the dissociation direction even in the absence of ppGpp and DksA (BOX 3b; FIG. 3a). Although ppGpp and DksA dramatically reduce the lifetimes of all promoter complexes that have been examined, they only inhibit transcription initiation from those promoters that form short-lived complexes with RNAP6,7,102,106. Presumably, ppGpp and DksA fail to inhibit output from promoters that make long-lived complexes because RNAP escapes to the elongation phase before promoter occupancy declines significantly.

Box 3. ppGpp and DksA inhibit some promoters and activate others.

Box 3

A kinetic framework for explaining both the direct negative and positive effects observed for ppGpp and DksA on transcription is shown in the figure. It is proposed that the binding of these factors to RNA polymerase (RNAP) lowers the free energy (ΔG) of an intermediate and/or transition state ([RP]) (dashed lines) between intermediates (labelled here as RP1 and RP2 on the pathway to open-complex formation)108. The consequence for transcription would depend on the intrinsic kinetic constants that are ultimately derived from the DNA sequence of an individual promoter. For example, at positively regulated amino acid promoters (see the figure, parta), the RP2 complex might be stable (have a lower free energy than RP1)108, which would make the promoter insensitive to the inhibitory effects of ppGpp and DksA. If there were also a high free-energy barrier between the two intermediates, ppGpp and DksA could lower this energy barrier, thereby decreasing the activation energy and increasing transcription initiation.

By contrast, at ribosomal RNA promoters (see the figure, partb), the activation energy for movement between the two free-energy states might be low and might not require ppGpp and DksA. However, if RP2 had a higher free energy than RP1, as for promoters that do not make stable open complexes at equilibrium, such asrrnBP1102,144, then the destabilizing effects of ppGpp and DksA would lower the barrier between the two states in the dissociation direction, thereby discouraging RP2 formation and inhibiting transcription. Thus, depending on the intrinsic kinetics of different promoters, ppGpp and DksA could regulate transcription on a global scale. The main kinetic properties that would define whether ppGpp and DksA have positive or negative effects on transcription would be the initial binding constant for RNAP and the dissociation rate of the competitor-resistant complex. The promoter sequences that define these kinetic properties are complex, which precludes identification of regulated promoters by sequence analysis alone6. Figure reproduced, with permission, from REF. 108 (2005) National

The features that contribute to the short half-life of the rRNA promoter complex with RNAP include a suboptimal −35 element, suboptimal spacing between the −10 and −35 hexamers, a suboptimal extended −10 element and a suboptimal discriminator sequence6. The UP element interaction with αCTD that recruits RNAP to the promoter does not contribute to the stability of the competitor-resistant complex. Therefore, the UP element has no effect on regulation by ppGpp and DksA6,7. Conversely, ppGpp and DksA probably have no effect on initial recruitment of RNAP to the rRNA promoter6,7,108. Separation of recruitment from the regulation functions of the promoter allows for finely tuned rates of ribosome synthesis even under conditions of high promoter occupancy.

High-resolution structures of ppGpp and DksA complexed with RNAP would greatly facilitate our understanding of their mechanisms of action. An X-ray crystal structure of ppGpp bound to RNAP holoenzyme from Thermus thermophilus113 revealed that ppGpp bound to a site that overlapped the so-called entry site for NTPs (the E site), as defined in the yeast RNA polymerase II elongation complex114. However, extensive biochemical and genetic analysis of the analogous site in E. coli RNAP, and the finding that ppGpp does not inhibit T. thermophilus RNAP in vitro or in vivo strongly suggest that this site is not responsible for the biologically significant effects of ppGpp on transcription initiation115,116.

RNAP preparations that lack the ω subunit are unable to respond to ppGpp in vitro117,118. The basis for the observed ω requirement in ppGpp function is not yet understood, but ω is not located near the E site of RNAP.

An X-ray crystal structure of E. coli DksA has also been solved107, although there is no RNAP–DksA co-crystal. The structure of DksA is reminiscent of that of GreA and GreB, bacterial transcription elongation factors that prevent transcription-elongation arrest and facilitate promoter escape119121. Similar to the Gre factors, DksA has a globular domain and a coiled-coil domain with an acidic tip and binds in the RNAP secondary channel107,122.

Exactly how DksA binding to RNAP leads to its effects on the lifetime of the promoter complex and on transcription initiation remains unclear. The proposal that the acidic residues at the tip of the DksA coiled-coil coordinate an Mg2+ ion bound to ppGpp107 requires re-evaluation if ppGpp functions elsewhere on RNAP115. Furthermore, this model does not address the mechanism that is responsible for the effects of DksA that are observed in the absence of ppGpp105,106,123,124 (discussed below).

Positive control of transcription by ppGpp and DksA

ppGpp and DksA work together to increase transcription from many promoters in response to nutrient limitation or other stresses. Some of these promoters are probably activated indirectly — that is, by the increase in RNAP concentration that results from the liberation of RNAP from rRNA operons (rRNA transcription represents the bulk of cellular transcription at high cellular growth rates)125,126. Consistent with the indirect (or passive) model, positively regulated promoters that have been found to be activated indirectly have low affinity for RNAP and require high concentrations of RNAP for transcription in vitro126. Furthermore, they make long-lived complexes with RNAP, which accounts for their insensitivity to direct inhibition by ppGpp and DksA102,126. ppGpp and DksA also seem to indirectly regulate some promoters that are transcribed by other RNAP holoenzymes: for example, Eσ54 and EσS (REFS 127,128). To explain the stimulatory effect of ppGpp on transcription from certain promoters that are dependent on alternative σ factors, it has also been suggested that ppGpp might directly facilitate the competition of alternative σ factors for core RNAP. However, to our knowledge, there is no evidence that ppGpp directly decreases the affinity of σ70 for core RNAP or directly increases the affinities of alternative σ factors for core RNAP (discussed in REF. 129).

Various promoters are also activated directly by ppGpp and DksA (BOX 3). These include the promoters for several amino acid biosynthesis operons108, promoters found in LEE (locus of enterocyte effacement) pathogenesis islands in pathogenic E. coli isolates109, some EσE-dependent promoters110 and the promoter for the global regulator Hfq111. Unsurprisingly, mutations in relA and spoT (which code for the enzymes that synthesize ppGpp) and mutations in dksA are pleiotropic.

NTP concentration and the control of transcription initiation

All promoters require higher concentrations of the iNTP than the subsequent NTPs, but certain promoters — for example, rRNA, some tRNA promoters7,103,130 and the fis promoter131 — require even higher concentrations of the iNTP than promoters in general. In extended stationary phase, depletion of NTPs therefore preferentially inhibits these promoters even though little or no ppGpp is present103,131. The rapid increase in NTP concentration is responsible for the increase in rRNA transcription that occurs when cells emerge from stationary phase103.

The mechanism by which changes in the iNTP concentration selectively affect rRNA synthesis has been debated. The transient stabilization of the intrinsically short-lived open complex by pairing of the first NTP (and subsequent NTPs) with the template strand may be sufficient to stimulate the initiation reaction by mass action7,102,130. Alternatively, it was proposed that binding of the first two NTPs results in a conformational change in RNAP and that this conformational change promotes transcript initiation independently of its role in stabilization of promoter–RNAP interactions132.

DksA amplifies the effects of low iNTP concentration on promoters that make intrinsically short-lived complexes with RNAP by increasing the probability of DNA-strand collapse before incorporation of the initial NTPs7,106. ppGpp-independent effects of DksA that have been observed on promoters in vivo123,124 might be attributed to the effects of promoter-complex lifetime on the iNTP concentration requirement. In addition, high DksA concentrations could increase the iNTP requirement by partially occluding the secondary channel, thereby reducing the rate of NTP diffusion to the active site7,105,106. It is also possible that there are times when regulation by DksA in vivo works independently of changes in ppGpp or NTP concentrations123,124.

The absence of a requirement for promoter sequence-specific DNA-binding proteins in regulation by ppGpp and iNTP concentrations suggests that the regulatory purposes of these factors evolved early in the history of life and therefore might occur in bacterial species that are evolutionarily distant from E. coli. Consistent with this hypothesis, changes in ppGpp and iNTP concentrations regulate rRNA promoters in B. subtilis133. However, ppGpp does not inhibit B. subtilis rRNA promoters directly. Rather, ppGpp reduces GTP levels and, because all B. subtilis rRNA promoters initiate with GTP, this reduction in GTP levels decreases rRNA promoter activity. Also consistent with this model is the fact that changing the identity of the base at position +1 in B. subtilis rRNA promoters results in a loss of regulation by ppGpp133. By contrast, E. coli rRNA transcription can begin with ATP, GTP or cytosine 5′-triphosphate (CTP) and still be regulated by ppGpp7,103,130. ppGpp probably reduces B. subtilis GTP pools in two ways: GTP is consumed in making ppGpp and ppGpp directly inhibits an enzyme in the GTP biosynthetic pathway97,133. Thus, E. coli and B. subtilis use the same signalling molecules to solve the same regulatory problem, but use different mechanisms.

Elegant systems also exist for the control of nucleotide biosynthesis operons in which NTP concentrations regulate gene expression. However, most of these mechanisms differ from those that control rRNA promoter activity. For example, in the E. coli pyrBI and carAB operons, an increase in the concentration of uridine 5′-triphosphate (UTP) results in reiterative transcription (which occurs when the initial transcript slips on the DNA template and non-coded nucleotides are added to the 3′ end of the RNA). High UTP levels lead to reiterative transcription and transcript release, which results in feedback inhibition of promoter activity134,135. pyrBI is also regulated at the level of transcription initiation by ppGpp and at the level of transcription elongation by UTP concentration-dependent attenuation104,134.

UTP concentrations also control transcription at the codBA and upp promoters. When UTP concentrations are high, transcripts initiate with ATP and form weak RNA–DNA hybrids that are more likely to undergo reiterative transcription, which leads to transcript release and decreased expression. When UTP concentrations are low, transcripts initiate with GTP and form stronger RNA–DNA hybrids that are less likely to undergo reiterative transcription136,137.

At the pyrC promoter, high CTP concentrations lead to initiation at a cytosine residue seven nucleotides from the end of the −10 hexamer, whereas low CTP concentration leads to initiation with GTP two nucleotides further downstream. Only the CTP-initiated transcript forms a hairpin that occludes ribosome binding138.

Changes in NTP concentration also seem to be crucial for regulation by the ~180-nucleotide 6S non-coding RNA139. Small RNAs regulate gene expression in all three biological kingdoms, but usually after the transcription-initiation step. However, 6S RNA is a promoter-specific inhibitor of transcription initiation that is found in a wide range of bacteria. 6S RNA binds stably to Eσ70 without binding to DNA, thereby competing with promoters for RNAP by mimicking a transcription bubble. Furthermore, Eσ70 can initiate RNA synthesis using 6S RNA as a template when the NTP concentration is high140. Because bound RNAP is released upon RNA synthesis, it has been proposed that regulation by 6S is achieved through changes in NTP concentration: when NTP concentrations are low (for example, during extended stationary phase), 6S traps RNAP, but when NTP concentrations rise during outgrowth from stationary phase, RNAP is released and is free to bind to promoters103,139,140.

Exactly how 6S RNA inhibits only some promoters is not fully understood141. Presumably, specific promoters have different affinities compared with 6S RNA for parts of σ70 (including region 4.2 (REF. 141)), and different rates of initiation of RNA synthesis from specific promoters and 6S RNA determine the magnitude of inhibition of specific promoters by 6S RNA. In any case, as in the examples described above, the intrinsic kinetic properties of the core promoter, as determined by its DNA sequence, determine its susceptibility to regulation by 6S RNA.

Osmoregulation

High external osmolarity increases cytoplasmic potassium glutamate and other organic anion concentrations by reducing the amount of cytoplasmic water. Bacteria have evolved elaborate strategies to deal with the high salt concentrations that can occur in their natural environments. At most promoters, the stability of open complexes, coupled with the accumulation of osmoprotectants and increased macromolecular crowding, buffer the RNAP–promoter interaction against the potential inhibitory effects of high external osmolarity (reviewed in REF. 142). However, rRNA promoters are dramatically inhibited when the salt concentration in E. coli cultures is increased143. The mechanism that is responsible has not been elucidated, but ppGpp does not seem to have a role in this response143. Because rRNA promoters are also especially sensitive to increased salt concentration in vitro144,145, it seems likely that the short lifetime and low level of RPO occupancy of the rRNA promoter complex results in its inhibition by high osmolarity in vivo. This inhibition of rRNA promoters by high osmolarity seems to be another situation in which the intrinsic kinetic characteristics of the rRNA promoter complex result in its specific regulation by environmental or nutritional signals.

Conclusions and future perspectives

The textbook description, in which bacterial promoter recognition resulted entirely from interactions between the −10 and −35 hexamers and the σ subunit of RNAP, has evolved in recent years with the discoveries of additional promoter elements and their interacting partners in RNAP. Likewise, the traditional view that the regulation of transcription initiation results primarily from activators and repressors that bind to specific DNA sequences has also been amended. We now know that regulators of transcription initiation can achieve specificity, in some cases, by binding directly to RNAP without binding to DNA and by exploiting the kinetic variation that ultimately results from differences in promoter sequences. We are only now beginning to develop an appreciation for the complexity of basal promoter activity and its role in transcription regulation in bacteria. The role of regulation of basal promoter activity should be a fertile area for research on the control of archaeal and eukaryotic gene expression in the future.

Acknowledgments

We thank colleagues in our laboratory and other laboratories who provided valuable input on sections of the text and figures, including C. Turnbough, R. Landick, R. Ebright, A. Hochschild and T. Record, and apologize to those investigators whose work has not been cited owing to space limitations. Our work is supported by the National Institutes of Health (grant R37 GM37048).

Glossary

Template strand

The strand of DNA that enters the active site of RNA polymerase and is used as a guide for RNA synthesis. +1 is defined as the position where the template strand pairs with the nucleoside 5′-triphosphate that forms the 5′ end of the transcript. The RNA transcript is the reverse complement of the template strand and has the same sequence as the non-template strand

Footprinting

A biochemical assay for detecting protein binding sites on DNA. A protein is allowed to bind to end-labelled DNA, the DNA is subjected to limited enzymatic or chemical nuclease cleavage and DNA fragments are separated by polyacrylamide electrophoresis under conditions that allow single-nucleotide resolution

Crosslinking

A biochemical technique for identifying interactions between macromolecules. Typically, covalent bonds between macromolecules are induced by ultraviolet or chemical exposure

Promoter clearance

The step in transcription in which RNAP breaks its interactions with the promoter and begins productive RNA synthesis

References

  • 1.Record MT, Jr, Reznikoff WS, Craig ML, McQuade KL, Schlax PJ. Escherichia coli and Salmonella. In: Neidhardt FC, editor. Cellular and Molecular Biology. ASM; Washington DC: 1996. pp. 792–820. [Google Scholar]
  • 2.Helmann JD, deHaseth PL. Protein–nucleic acid interactions during open complex formation investigated by systematic alteration of the protein and DNA binding partners. Biochemistry. 1999;38:5959–5967. doi: 10.1021/bi990206g. [DOI] [PubMed] [Google Scholar]
  • 3.Browning DF, Busby SJ. The regulation of bacterial transcription initiation. Nature Rev Microbiol. 2004;2:57–65. doi: 10.1038/nrmicro787. [DOI] [PubMed] [Google Scholar]
  • 4.Ross W, et al. A third recognition element in bacterial promoters: DNA binding by the alpha subunit of RNA polymerase. Science. 1993;262:1407–1413. doi: 10.1126/science.8248780. [DOI] [PubMed] [Google Scholar]
  • 5.Barne KA, Bown JA, Busby SJ, Minchin SD. Region 2.5 of the Escherichia coli RNA polymerase σ70 subunit is responsible for the recognition of the ‘extended-10’ motif at promoters. EMBO J. 1997;16:4034–4040. doi: 10.1093/emboj/16.13.4034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Haugen SP, et al. rRNA promoter regulation by nonoptimal binding of σ region 1.2: an additional recognition element for RNA polymerase. Cell. 2006;125:1069–1082. doi: 10.1016/j.cell.2006.04.034. [DOI] [PubMed] [Google Scholar]
  • 7.Paul BJ, Ross W, Gaal T, Gourse RL. rRNA transcription in Escherichia coli. Annu Rev Genet. 2004;38:749–770. doi: 10.1146/annurev.genet.38.072902.091347. [DOI] [PubMed] [Google Scholar]
  • 8.Minakhin L, et al. Bacterial RNA polymerase subunit ω and eukaryotic RNA polymerase subunit RPB6 are sequence, structural, and functional homologs and promote RNA polymerase assembly. Proc Natl Acad Sci USA. 2001;98:892–897. doi: 10.1073/pnas.98.3.892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Zhang G, et al. Crystal structure of Thermus aquaticus core RNA polymerase at 3.3 Å resolution. Cell. 1999;98:811–824. doi: 10.1016/s0092-8674(00)81515-9. [DOI] [PubMed] [Google Scholar]
  • 10.Hirata A, Klein BJ, Murakami KS. The X-ray crystal structure of RNA polymerase from Archaea. Nature. 2008;451:851–854. doi: 10.1038/nature06530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Cramer P, Bushnell DA, Kornberg RD. Structural basis of transcription: RNA polymerase II at 2.8 Å resolution. Science. 2001;292:1863–1876. doi: 10.1126/science.1059493. [DOI] [PubMed] [Google Scholar]
  • 12.Murakami KS, Masuda S, Darst SA. Structural basis of transcription initiation: RNA polymerase holoenzyme at 4 Å resolution. Science. 2002;296:1280–1284. doi: 10.1126/science.1069594. [DOI] [PubMed] [Google Scholar]
  • 13.Vassylyev DG, et al. Crystal structure of a bacterial RNA polymerase holoenzyme at 2.6 Å resolution. Nature. 2002;417:712–719. doi: 10.1038/nature752. [DOI] [PubMed] [Google Scholar]
  • 14.Murakami KS, Masuda S, Campbell EA, Muzzin O, Darst SA. Structural basis of transcription initiation: an RNA polymerase holoenzyme–DNA complex. Science. 2002;296:1285–1290. doi: 10.1126/science.1069595. [DOI] [PubMed] [Google Scholar]
  • 15.Murakami KS, Darst SA. Bacterial RNA polymerases: the wholo story. Curr Opin Struct Biol. 2003;13:31–39. doi: 10.1016/s0959-440x(02)00005-2. [DOI] [PubMed] [Google Scholar]
  • 16.Landick R. NTP-entry routes in multi-subunit RNA polymerases. Trends Biochem Sci. 2005;30:651–654. doi: 10.1016/j.tibs.2005.10.001. [DOI] [PubMed] [Google Scholar]
  • 17.Korzheva N, et al. A structural model of transcription elongation. Science. 2000;289:619–625. doi: 10.1126/science.289.5479.619. [DOI] [PubMed] [Google Scholar]
  • 18.Vassylyev DG, Vassylyeva MN, Perederina A, Tahirov TH, Artsimovitch I. Structural basis for transcription elongation by bacterial RNA polymerase. Nature. 2007;448:157–162. doi: 10.1038/nature05932. [DOI] [PubMed] [Google Scholar]
  • 19.Lawson CL, et al. Catabolite activator protein: DNA binding and transcription activation. Curr Opin Struct Biol. 2004;14:10–20. doi: 10.1016/j.sbi.2004.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Blatter EE, Ross W, Tang H, Gourse RL, Ebright RH. 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]
  • 21.Mekler V, et al. Structural organization of bacterial RNA polymerase holoenzyme and the RNA polymerase–promoter open complex. Cell. 2002;108:599–614. doi: 10.1016/s0092-8674(02)00667-0. [DOI] [PubMed] [Google Scholar]
  • 22.Saecker RM, et al. Kinetic studies and structural models of the association of E. coli σ70 RNA polymerase with the λ PR promoter: large scale conformational changes in forming the kinetically significant intermediates. J Mol Biol. 2002;319:649–671. doi: 10.1016/S0022-2836(02)00293-0. [DOI] [PubMed] [Google Scholar]
  • 23.Li XY, McClure WR. Characterization of the closed complex intermediate formed during transcription initiation by Escherichia coli RNA polymerase. J Biol Chem. 1998;273:23549–23557. doi: 10.1074/jbc.273.36.23549. [DOI] [PubMed] [Google Scholar]
  • 24.Cook VM, deHaseth PL. Strand opening-deficient E. coli RNA polymerase facilitates investigation of closed complexes with promoter DNA: effects of DNA sequence and temperature. J Biol Chem. 2007;282:21319–21326. doi: 10.1074/jbc.M702232200. [DOI] [PubMed] [Google Scholar]
  • 25.Suh WC, Ross W, Record MT., Jr Two open complexes and a requirement for Mg2+ to open the lambda PR transcription start site. Science. 1993;259:358–361. doi: 10.1126/science.8420002. [DOI] [PubMed] [Google Scholar]
  • 26.Naryshkin N, Revyakin A, Kim Y, Mekler V, Ebright RH. Structural organization of the RNA polymerase–promoter open complex. Cell. 2000;101:601–611. doi: 10.1016/s0092-8674(00)80872-7. [DOI] [PubMed] [Google Scholar]
  • 27.Kontur WS, Saecker RM, Capp MW, Record MT., Jr Late steps in the formation of E.coli RNA polymerase — λPR promoter open complexes: characterization of conformational changes by rapid [perturbant] upshift experiments. J Mol Biol. 2008;376:1034–1047. doi: 10.1016/j.jmb.2007.11.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Davis CA, Bingman CA, Landick R, Record MT, Jr, Saecker RM. Real-time footprinting of DNA in the first kinetically significant intermediate in open complex formation by Escherichia coli RNA polymerase. Proc Natl Acad Sci USA. 2007;104:7833–7838. doi: 10.1073/pnas.0609888104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sclavi B, et al. Real-time characterization of intermediates in the pathway to open complex formation by Escherichia coli RNA polymerase at the T7A1 promoter. Proc Natl Acad Sci USA. 2005;102:4706–4711. doi: 10.1073/pnas.0408218102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Geszvain KM, Landick R. In: The Bacterial Chromosome. Higgins NP, editor. ASM; Washington DC: 2005. pp. 283–296. [Google Scholar]
  • 31.Haugen S, Ross W, Manrique M, Gourse RL. Fine structure of the promoter–σ region 1.2 interaction. Proc Natl Acad Sci USA. 2008;105:3292–3297. doi: 10.1073/pnas.0709513105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Feklistov A, et al. A basal promoter element recognized by free RNA polymerase σ subunit determines promoter recognition by RNA polymerase holoenzyme. Mol Cell. 2006;23:97–107. doi: 10.1016/j.molcel.2006.06.010. [DOI] [PubMed] [Google Scholar]
  • 33.Young BA, Gruber TM, Gross CA. Views of transcription initiation. Cell. 2002;109:417–420. doi: 10.1016/s0092-8674(02)00752-3. [DOI] [PubMed] [Google Scholar]
  • 34.Revyakin A, Liu C, Ebright RH, Strick TR. Abortive initiation and productive initiation by RNA polymerase involve DNA scrunching. Science. 2006;314:1139–1143. doi: 10.1126/science.1131398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kapanidis AN, et al. Initial transcription by RNA polymerase proceeds through a DNA-scrunching mechanism. Science. 2006;314:1144–1147. doi: 10.1126/science.1131399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Nickels BE, et al. The interaction between σ70 and the β-flap of Escherichia coli RNA polymerase inhibits extension of nascent RNA during early elongation. Proc Natl Acad Sci USA. 2005;102:4488–4493. doi: 10.1073/pnas.0409850102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Gourse RL, Ross W, Gaal T. UPs and downs in bacterial transcription initiation: the role of the alpha subunit of RNA polymerase in promoter recognition. Mol Microbiol. 2000;37:687–695. doi: 10.1046/j.1365-2958.2000.01972.x. [DOI] [PubMed] [Google Scholar]
  • 38.Ross W, Ernst A, Gourse RL. Fine structure of E. coli RNA polymerase–promoter interactions: α subunit binding to the UP element minor groove. Genes Dev. 2001;15:491–506. doi: 10.1101/gad.870001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lee DJ, Busby SJ, Lloyd GS. Exploitation of a chemical nuclease to investigate the location and orientation of the Escherichia coli RNA polymerase α subunit C-terminal domains at simple promoters that are activated by cyclic AMP receptor protein. J Biol Chem. 2003;278:52944–52952. doi: 10.1074/jbc.M308300200. [DOI] [PubMed] [Google Scholar]
  • 40.Benoff B, et al. Structural basis of transcription activation: the CAP-αCTD–DNA complex. Science. 2002;297:1562–1566. doi: 10.1126/science.1076376. [DOI] [PubMed] [Google Scholar]
  • 41.Ross W, Gourse RL. Sequence-independent upstream DNA αCTD interactions strongly stimulate Escherichia coli RNA polymerase-lacUV5 promoter association. Proc Natl Acad Sci USA. 2005;102:291–296. doi: 10.1073/pnas.0405814102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Strainic MG, Jr, Sullivan JJ, Velevis A, deHaseth PL. Promoter recognition by Escherichia coli RNA polymerase: effects of the UP element on open complex formation and promoter clearance. Biochemistry. 1998;37:18074–18080. doi: 10.1021/bi9813431. [DOI] [PubMed] [Google Scholar]
  • 43.Davis CA, Capp MW, Record MT, Jr, Saecker RM. The effects of upstream DNA on open complex formation by Escherichia coli RNA polymerase. Proc Natl Acad Sci USA. 2005;102:285–290. doi: 10.1073/pnas.0405779102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Campbell EA, et al. Structure of the bacterial RNA polymerase promoter specificity σ subunit. Mol Cell. 2002;9:527–539. doi: 10.1016/s1097-2765(02)00470-7. [DOI] [PubMed] [Google Scholar]
  • 45.Jain D, Nickels BE, Sun L, Hochschild A, Darst SA. Structure of a ternary transcription activation complex. Mol Cell. 2004;13:45–53. doi: 10.1016/s1097-2765(03)00483-0. [DOI] [PubMed] [Google Scholar]
  • 46.Eichenberger P, Dethiollaz S, Buc H, Geiselmann J. Structural kinetics of transcription activation at the malT promoter of Escherichia coli by UV laser footprinting. Proc Natl Acad Sci USA. 1997;94:9022–9027. doi: 10.1073/pnas.94.17.9022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Buckle M, Pemberton IK, Jacquet MA, Buc H. The kinetics of sigma subunit directed promoter recognition by E. coli RNA polymerase. J Mol Biol. 1999;285:955–964. doi: 10.1006/jmbi.1998.2391. [DOI] [PubMed] [Google Scholar]
  • 48.Ross W, Schneider DA, Paul BJ, Mertens A, Gourse RL. An intersubunit contact stimulating transcription initiation by E. coli RNA polymerase: interaction of the α C-terminal domain and σ region 4. Genes Dev. 2003;17:1293–1307. doi: 10.1101/gad.1079403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Chen H, Tang H, Ebright RH. Functional interaction between RNA polymerase α subunit C-terminal domain and σ70 in UP-element- and activator-dependent transcription. Mol Cell. 2003;11:1621–1633. doi: 10.1016/s1097-2765(03)00201-6. [DOI] [PubMed] [Google Scholar]
  • 50.Mitchell JE, Zheng D, Busby SJ, Minchin SD. Identification and analysis of ‘extended −10’ promoters in Escherichia coli. Nucleic Acids Res. 2003;31:4689–4695. doi: 10.1093/nar/gkg694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Thouvenot B, Charpentier B, Branlant C. The strong efficiency of the Escherichia coli gapA P1 promoter depends on a complex combination of functional determinants. Biochem J. 2004;383:371–382. doi: 10.1042/BJ20040792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Hook-Barnard I, Johnson XB, Hinton DM. Escherichia coli RNA polymerase recognition of a σ70-dependent promoter requiring a −35 DNA element and an extended −10 TGn Motif. J Bacteriol. 2006;188:8352–8359. doi: 10.1128/JB.00853-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Camacho A, Salas M. Effect of mutations in the “extended −10” motif of three Bacillus subtilis σ A–RNA polymerase-dependent promoters. J Mol Biol. 1999;286:683–693. doi: 10.1006/jmbi.1998.2526. [DOI] [PubMed] [Google Scholar]
  • 54.Voskuil MI, Chambliss GH. The TRTGn motif stabilizes the transcription initiation open complex. J Mol Biol. 2002;322:521–532. doi: 10.1016/s0022-2836(02)00802-1. [DOI] [PubMed] [Google Scholar]
  • 55.Hengge-Aronis R. Stationary phase gene regulation: what makes an Escherichia coli promoter σS-selective? Curr Opin Microbiol. 2002;5:591–595. doi: 10.1016/s1369-5274(02)00372-7. [DOI] [PubMed] [Google Scholar]
  • 56.Matlock DL, Heyduk T. Sequence determinants for the recognition of the fork junction DNA containing the −10 region of promoter DNA by E. coli RNA polymerase. Biochemistry. 2000;39:12274–12283. doi: 10.1021/bi001433h. [DOI] [PubMed] [Google Scholar]
  • 57.Young BA, Gruber TM, Gross CA. Minimal machinery of RNA polymerase holoenzyme sufficient for promoter melting. Science. 2004;303:1382–1384. doi: 10.1126/science.1092462. [DOI] [PubMed] [Google Scholar]
  • 58.Roberts CW, Roberts JW. Base-specific recognition of the nontemplate strand of promoter DNA by E. coli RNA polymerase. Cell. 1996;86:495–501. doi: 10.1016/s0092-8674(00)80122-1. [DOI] [PubMed] [Google Scholar]
  • 59.Marr MT, Roberts JW. Promoter recognition as measured by binding of polymerase to nontemplate strand oligonucleotide. Science. 1997;276:1258–1260. doi: 10.1126/science.276.5316.1258. [DOI] [PubMed] [Google Scholar]
  • 60.Fenton MS, Lee SJ, Gralla JD. Escherichia coli promoter opening and −10 recognition: mutational analysis of σ70. EMBO J. 2000;19:1130–1137. doi: 10.1093/emboj/19.5.1130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Qiu J, Helmann JD. Adenines at −11, −9 and −8 play a key role in the binding of Bacillus subtilis EσA RNA polymerase to −10 region single-stranded DNA. Nucleic Acids Res. 1999;27:4541–4546. doi: 10.1093/nar/27.23.4541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Tomsic M, et al. Different roles for basic and aromatic amino acids in conserved region 2 of Escherichia coli σ70 in the nucleation and maintenance of the single-stranded DNA bubble in open RNA polymerase–promoter complexes. J Biol Chem. 2001;276:31891–31896. doi: 10.1074/jbc.M105027200. [DOI] [PubMed] [Google Scholar]
  • 63.Lim HM, Lee HJ, Roy S, Adhya S. A “master” in base unpairing during isomerization of a promoter upon RNA polymerase binding. Proc Natl Acad Sci USA. 2001;98:14849–14852. doi: 10.1073/pnas.261517398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Tsujikawa L, Strainic MG, Watrob H, Barkley MD, DeHaseth PL. RNA polymerase alters the mobility of an A-residue crucial to polymerase-induced melting of promoter DNA. Biochemistry. 2002;41:15334–15341. doi: 10.1021/bi026539m. [DOI] [PubMed] [Google Scholar]
  • 65.Lee HJ, Lim HM, Adhya S. An unsubstituted C2 hydrogen of adenine is critical and sufficient at the −11 position of a promoter to signal base pair deformation. J Biol Chem. 2004;279:16899–16902. doi: 10.1074/jbc.C400054200. [DOI] [PubMed] [Google Scholar]
  • 66.Heyduk E, Kuznedelov K, Severinov K, Heyduk T. A consensus adenine at position −11 of the nontemplate strand of bacterial promoter is important for nucleation of promoter melting. J Biol Chem. 2006;281:12362–12369. doi: 10.1074/jbc.M601364200. [DOI] [PubMed] [Google Scholar]
  • 67.McClure WR, Hawley DK, Youderian P, Susskind MM. DNA determinants of promoter selectivity in Escherichia coli. Cold Spring Harb Symp Quant Biol. 1983;47 (Pt 1):477–481. doi: 10.1101/sqb.1983.047.01.057. [DOI] [PubMed] [Google Scholar]
  • 68.Travers AA. Promoter sequence for stringent control of bacterial ribonucleic acid synthesis. J Bacteriol. 1980;141:973–976. doi: 10.1128/jb.141.2.973-976.1980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Travers AA, Burgess RR. Cyclic re-use of the RNA polymerase sigma factor. Nature. 1969;222:537–540. doi: 10.1038/222537a0. [DOI] [PubMed] [Google Scholar]
  • 70.Bar-Nahum G, Nudler E. Isolation and characterization of σ70-retaining transcription elongation complexes from Escherichia coli. Cell. 2001;106:443–451. doi: 10.1016/s0092-8674(01)00461-5. [DOI] [PubMed] [Google Scholar]
  • 71.Mukhopadhyay J, et al. Translocation of σ70 with RNA polymerase during transcription: fluorescence resonance energy transfer assay for movement relative to DNA. Cell. 2001;106:453–463. doi: 10.1016/s0092-8674(01)00464-0. [DOI] [PubMed] [Google Scholar]
  • 72.Mooney RA, Darst SA, Landick R. Sigma and RNA polymerase: an on-again, off-again relationship? Mol Cell. 2005;20:335–345. doi: 10.1016/j.molcel.2005.10.015. [DOI] [PubMed] [Google Scholar]
  • 73.Raffaelle M, Kanin EI, Vogt J, Burgess RR, Ansari AZ. Holoenzyme switching and stochastic release of sigma factors from RNA polymerase in vivo. Mol Cell. 2005;20:357–366. doi: 10.1016/j.molcel.2005.10.011. [DOI] [PubMed] [Google Scholar]
  • 74.Reppas NB, Wade JT, Church GM, Struhl K. The transition between transcriptional initiation and elongation in E. coli is highly variable and often rate limiting. Mol Cell. 2006;24:747–757. doi: 10.1016/j.molcel.2006.10.030. [DOI] [PubMed] [Google Scholar]
  • 75.Shimamoto N, Kamigochi T, Utiyama H. Release of the sigma subunit of Escherichia coli DNA-dependent RNA polymerase depends mainly on time elapsed after the start of initiation, not on length of product RNA. J Biol Chem. 1986;261:11859–11865. [PubMed] [Google Scholar]
  • 76.Sharp MM, et al. The interface of σ with core RNA polymerase is extensive, conserved, and functionally specialized. Genes Dev. 1999;13:3015–3026. doi: 10.1101/gad.13.22.3015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Gruber TM, et al. Binding of the initiation factor σ70 to core RNA polymerase is a multistep process. Mol Cell. 2001;8:21–31. doi: 10.1016/s1097-2765(01)00292-1. [DOI] [PubMed] [Google Scholar]
  • 78.Young BA, et al. A coiled-coil from the RNA polymerase β′ subunit allosterically induces selective nontemplate strand binding by σ70. Cell. 2001;105:935–944. doi: 10.1016/s0092-8674(01)00398-1. [DOI] [PubMed] [Google Scholar]
  • 79.Leibman M, Hochschild A. A σ-core interaction of the RNA polymerase holoenzyme that enhances promoter escape. EMBO J. 2007;26:1579–1590. doi: 10.1038/sj.emboj.7601612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Quinones M, Kimsey HH, Ross W, Gourse RL, Waldor MK. LexA represses CTXφ transcription by blocking access of the α C-terminal domain of RNA polymerase to promoter DNA. J Biol Chem. 2006;281:39407–39412. doi: 10.1074/jbc.M609694200. [DOI] [PubMed] [Google Scholar]
  • 81.Ansari AZ, Bradner JE, O’Halloran TV. DNA-bend modulation in a repressor-to-activator switching mechanism. Nature. 1995;374:371–375. doi: 10.1038/374370a0. [DOI] [PubMed] [Google Scholar]
  • 82.Choy HE, et al. Repression and activation of promoter-bound RNA polymerase activity by Gal repressor. J Mol Biol. 1997;272:293–300. doi: 10.1006/jmbi.1997.1221. [DOI] [PubMed] [Google Scholar]
  • 83.Tagami H, Aiba H. An inactive open complex mediated by an UP element at Escherichia coli promoters. Proc Natl Acad Sci USA. 1999;96:7202–7207. doi: 10.1073/pnas.96.13.7202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Beck LL, Smith TG, Hoover TR. Look, no hands! Unconventional transcriptional activators in bacteria. Trends Microbiol. 2007;15:530–537. doi: 10.1016/j.tim.2007.09.008. [DOI] [PubMed] [Google Scholar]
  • 85.Helmann JD. Anti-sigma factors. Curr Opin Microbiol. 1999;2:135–141. doi: 10.1016/S1369-5274(99)80024-1. [DOI] [PubMed] [Google Scholar]
  • 86.Campbell EA, et al. A conserved structural module regulates transcriptional responses to diverse stress signals in bacteria. Mol Cell. 2007;27:793–805. doi: 10.1016/j.molcel.2007.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Gregory BD, et al. A regulator that inhibits transcription by targeting an intersubunit interaction of the RNA polymerase holoenzyme. Proc Natl Acad Sci USA. 2004;101:4554–4559. doi: 10.1073/pnas.0400923101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Lambert LJ, Wei Y, Schirf V, Demeler B, Werner MH. T4 AsiA blocks DNA recognition by remodeling σ70 region 4. EMBO J. 2004;23:2952–2962. doi: 10.1038/sj.emboj.7600312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Zuber P. Spx–RNA polymerase interaction and global transcriptional control during oxidative stress. J Bacteriol. 2004;186:1911–1918. doi: 10.1128/JB.186.7.1911-1918.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Nakano S, Erwin KN, Ralle M, Zuber P. Redox-sensitive transcriptional control by a thiol/disulphide switch in the global regulator, Spx. Mol Microbiol. 2005;55:498–510. doi: 10.1111/j.1365-2958.2004.04395.x. [DOI] [PubMed] [Google Scholar]
  • 91.Pratt LA, Silhavy TJ. Crl stimulates RpoS activity during stationary phase. Mol Microbiol. 1998;29:1225–1236. doi: 10.1046/j.1365-2958.1998.01007.x. [DOI] [PubMed] [Google Scholar]
  • 92.Typas A, Barembruch C, Possling A, Hengge R. Stationary phase reorganisation of the Escherichia coli transcription machinery by Crl protein, a fine-tuner of σS activity and levels. EMBO J. 2007;26:1569–1578. doi: 10.1038/sj.emboj.7601629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Gaal T, Mandel MJ, Silhavy TJ, Gourse RL. Crl facilitates RNA polymerase holoenzyme formation. J Bacteriol. 2006;188:7966–7970. doi: 10.1128/JB.01266-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Bougdour A, Lelong C, Geiselmann J. Crl, a low temperature-induced protein in Escherichia coli that binds directly to the stationary phase σ subunit of RNA polymerase. J Biol Chem. 2004;279:19540–19550. doi: 10.1074/jbc.M314145200. [DOI] [PubMed] [Google Scholar]
  • 95.Robbe-Saule V, Lopes MD, Kolb A, Norel F. Physiological effects of Crl in Salmonella are modulated by σS level and promoter specificity. J Bacteriol. 2007;189:2976–2987. doi: 10.1128/JB.01919-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Robbe-Saule V, et al. Crl activates transcription initiation of RpoS-regulated genes involved in the multicellular behavior of Salmonella enterica serovar Typhimurium. J Bacteriol. 2006;188:3983–3994. doi: 10.1128/JB.00033-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Potrykus K, Cashel M. (p)ppGpp: still Magical? Annu Rev Microbiol. 2008;62:35–51. doi: 10.1146/annurev.micro.62.081307.162903. [DOI] [PubMed] [Google Scholar]
  • 98.Takahashi K, Kasai K, Ochi K. Identification of the bacterial alarmone guanosine 5′-diphosphate 3′-diphosphate (ppGpp) in plants. Proc Natl Acad Sci USA. 2004;101:4320–4324. doi: 10.1073/pnas.0308555101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Traxler MF, Chang DE, Conway T. Guanosine 3′,5′-bispyrophosphate coordinates global gene expression during glucose-lactose diauxie in Escherichia coli. Proc Natl Acad Sci USA. 2006;103:2374–2379. doi: 10.1073/pnas.0510995103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Durfee T, Hansen AM, Zhi H, Blattner FR, Jin DJ. Transcription profiling of the stringent response in Escherichia coli. J Bacteriol. 2008;190:1084–1096. doi: 10.1128/JB.01092-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Eymann C, Homuth G, Scharf C, Hecker M. Bacillus subtilis functional genomics: global characterization of the stringent response by proteome and transcriptome analysis. J Bacteriol. 2002;184:2500–2520. doi: 10.1128/JB.184.9.2500-2520.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Barker MM, Gaal T, Josaitis CA, Gourse RL. Mechanism of regulation of transcription initiation by ppGpp. I Effects of ppGpp on transcription initiation in vivo and in vitro. J Mol Biol. 2001;305:673–688. doi: 10.1006/jmbi.2000.4327. [DOI] [PubMed] [Google Scholar]
  • 103.Murray HD, Schneider DA, Gourse RL. Control of rRNA expression by small molecules is dynamic and nonredundant. Mol Cell. 2003;12:125–134. doi: 10.1016/s1097-2765(03)00266-1. [DOI] [PubMed] [Google Scholar]
  • 104.Donahue JP, Turnbough CL., Jr Characterization of transcriptional initiation from promoters P1 and P2 of the pyrBI operon of Escherichia coli K12. J Biol Chem. 1990;265:19091–19099. [PubMed] [Google Scholar]
  • 105.Mallik P, Paul BJ, Rutherford ST, Gourse RL, Osuna R. DksA is required for growth phase-dependent regulation, growth rate-dependent control, and stringent control of fis expression in Escherichia coli. J Bacteriol. 2006;188:5775–5782. doi: 10.1128/JB.00276-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Paul BJ, et al. DksA: a critical component of the transcription initiation machinery that potentiates the regulation of rRNA promoters by ppGpp and the initiating NTP. Cell. 2004;118:311–322. doi: 10.1016/j.cell.2004.07.009. [DOI] [PubMed] [Google Scholar]
  • 107.Perederina A, et al. Regulation through the secondary channel — structural framework for ppGpp–DksA synergism during transcription. Cell. 2004;118:297–309. doi: 10.1016/j.cell.2004.06.030. [DOI] [PubMed] [Google Scholar]
  • 108.Paul BJ, Berkmen MB, Gourse RL. DksA potentiates direct activation of amino acid promoters by ppGpp. Proc Natl Acad Sci USA. 2005;102:7823–7828. doi: 10.1073/pnas.0501170102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Nakanishi N, et al. ppGpp with DksA controls gene expression in the locus of enterocyte effacement (LEE) pathogenicity island of enterohaemorrhagic Escherichia coli through activation of two virulence regulatory genes. Mol Microbiol. 2006;61:194–205. doi: 10.1111/j.1365-2958.2006.05217.x. [DOI] [PubMed] [Google Scholar]
  • 110.Costanzo A, et al. Mechanism of regulation of the extracytoplasmic stress factor σE in Escherichia coli by DksA and the alarmone ppGpp. Mol Microbiol. 2008;67:619–632. doi: 10.1111/j.1365-2958.2007.06072.x. [DOI] [PubMed] [Google Scholar]
  • 111.Sharma AK, Payne SM. Induction of expression of hfq by DksA is essential for Shigella flexneri virulence. Mol Microbiol. 2006;62:469–479. doi: 10.1111/j.1365-2958.2006.05376.x. [DOI] [PubMed] [Google Scholar]
  • 112.Choy HE. The study of guanosine 5′-diphosphate 3′-diphosphate-mediated transcription regulation in vitro using a coupled transcription–translation system. J Biol Chem. 2000;275:6783–6789. doi: 10.1074/jbc.275.10.6783. [DOI] [PubMed] [Google Scholar]
  • 113.Artsimovitch I, et al. Structural basis for transcription regulation by alarmone ppGpp. Cell. 2004;117:299–310. doi: 10.1016/s0092-8674(04)00401-5. [DOI] [PubMed] [Google Scholar]
  • 114.Westover KD, Bushnell DA, Kornberg RD. Structural basis of transcription: nucleotide selection by rotation in the RNA polymerase II active center. Cell. 2004;119:481–489. doi: 10.1016/j.cell.2004.10.016. [DOI] [PubMed] [Google Scholar]
  • 115.Vrentas CE, et al. Still looking for the magic spot: the crystallographically-defined binding site for ppGpp on RNA polymerase is unlikely to be responsible for rRNA transcription regulation. J Mol Biol. 2008;377:551–564. doi: 10.1016/j.jmb.2008.01.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Kasai K, et al. Physiological analysis of the stringent response elicited in an extreme thermophilic bacterium, Thermus thermophilus. J Bacteriol. 2006;188:7111–7122. doi: 10.1128/JB.00574-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Igarashi K, Fujita N, Ishihama A. Promoter selectivity of Escherichia coli RNA polymerase: omega factor is responsible for the ppGpp sensitivity. Nucleic Acids Res. 1989;17:8755–8765. doi: 10.1093/nar/17.21.8755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Vrentas CE, Gaal T, Ross W, Ebright RH, Gourse RL. Response of RNA polymerase to ppGpp: requirement for the ω subunit and relief of this requirement by DksA. Genes Dev. 2005;19:2378–2387. doi: 10.1101/gad.1340305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Opalka N, et al. Structure and function of the transcription elongation factor GreB bound to bacterial RNA polymerase. Cell. 2003;114:335–345. doi: 10.1016/s0092-8674(03)00600-7. [DOI] [PubMed] [Google Scholar]
  • 120.Laptenko O, Lee J, Lomakin I, Borukhov S. Transcript cleavage factors GreA and GreB act as transient catalytic components of RNA polymerase. EMBO J. 2003;22:6322–6334. doi: 10.1093/emboj/cdg610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Stepanova E, et al. Analysis of promoter targets for Escherichia coli transcription elongation factor GreA in vivo and in vitro. J Bacteriol. 2007;189:8772–8785. doi: 10.1128/JB.00911-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Rutherford ST, et al. Effects of DksA, GreA, and GreB on transcription initiation: insights into the mechanisms of factors that bind in the secondary channel of RNA polymerase. J Mol Biol. 2007;366:1243–1257. doi: 10.1016/j.jmb.2006.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Magnusson LU, Gummesson B, Joksimovic P, Farewell A, Nystrom T. Identical, independent, and opposing roles of ppGpp and DksA in Escherichia coli. J Bacteriol. 2007;189:5193–5202. doi: 10.1128/JB.00330-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Aberg A, Shingler V, Balsalobre C. Regulation of the fimB promoter: a case of differential regulation by ppGpp and DksA in vivo. Mol Microbiol. 2008;67:1223–1241. doi: 10.1111/j.1365-2958.2008.06115.x. [DOI] [PubMed] [Google Scholar]
  • 125.Zhou YN, Jin DJ. The rpoB mutants destabilizing initiation complexes at stringently controlled promoters behave like “stringent” RNA polymerases in Escherichia coli. Proc Natl Acad Sci USA. 1998;95:2908–2913. doi: 10.1073/pnas.95.6.2908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Barker MM, Gaal T, Gourse RL. Mechanism of regulation of transcription initiation by ppGpp. II Models for positive control based on properties of RNAP mutants and competition for RNAP. J Mol Biol. 2001;305:689–702. doi: 10.1006/jmbi.2000.4328. [DOI] [PubMed] [Google Scholar]
  • 127.Bernardo LM, Johansson LU, Solera D, Skarfstad E, Shingler V. The guanosine tetraphosphate (ppGpp) alarmone, DksA and promoter affinity for RNA polymerase in regulation of σ22-dependent transcription. Mol Microbiol. 2006;60:749–764. doi: 10.1111/j.1365-2958.2006.05129.x. [DOI] [PubMed] [Google Scholar]
  • 128.Magnusson LU, Farewell A, Nystrom T. ppGpp: a global regulator in Escherichia coli. Trends Microbiol. 2005;13:236–242. doi: 10.1016/j.tim.2005.03.008. [DOI] [PubMed] [Google Scholar]
  • 129.Gourse RL, Ross W, Rutherford ST. General pathway for turning on promoters transcribed by RNA polymerases containing alternative σ factors. J Bacteriol. 2006;188:4589–4591. doi: 10.1128/JB.00499-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Gaal T, Bartlett MS, Ross W, Turnbough CL, Jr, Gourse RL. Transcription regulation by initiating NTP concentration: rRNA synthesis in bacteria. Science. 1997;278:2092–2097. doi: 10.1126/science.278.5346.2092. [DOI] [PubMed] [Google Scholar]
  • 131.Walker KA, Mallik P, Pratt TS, Osuna R. The Escherichia coli fis promoter is regulated by changes in the levels of its transcription initiation nucleotide CTP. J Biol Chem. 2004;279:50818–50828. doi: 10.1074/jbc.M406285200. [DOI] [PubMed] [Google Scholar]
  • 132.Lew CM, Gralla JD. Nucleotide-dependent isomerization of Escherichia coli RNA polymerase. Biochemistry. 2004;43:12660–12666. doi: 10.1021/bi0492814. [DOI] [PubMed] [Google Scholar]
  • 133.Krasny L, Gourse RL. An alternative strategy for bacterial ribosome synthesis: Bacillus subtilis rRNA transcription regulation. EMBO J. 2004;23:4473–4483. doi: 10.1038/sj.emboj.7600423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Liu C, Heath LS, Turnbough CL., Jr Regulation of pyrBI operon expression in Escherichia coli by UTP-sensitive reiterative RNA synthesis during transcriptional initiation. Genes Dev. 1994;8:2904–2912. doi: 10.1101/gad.8.23.2904. [DOI] [PubMed] [Google Scholar]
  • 135.Han X, Turnbough CL., Jr Regulation of carAB expression in Escherichia coli occurs in part through UTP-sensitive reiterative transcription. J Bacteriol. 1998;180:705–713. doi: 10.1128/jb.180.3.705-713.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Qi F, Turnbough CL., Jr Regulation of codBA operon expression in Escherichia coli by UTP-dependent reiterative transcription and UTP-sensitive transcriptional start site switching. J Mol Biol. 1995;254:552–565. doi: 10.1006/jmbi.1995.0638. [DOI] [PubMed] [Google Scholar]
  • 137.Tu AH, Turnbough CL., Jr Regulation of upp expression in Escherichia coli by UTP-sensitive selection of transcriptional start sites coupled with UTP-dependent reiterative transcription. J Bacteriol. 1997;179:6665–6673. doi: 10.1128/jb.179.21.6665-6673.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Wilson HR, Archer CD, Liu JK, Turnbough CL., Jr Translational control of pyrC expression mediated by nucleotide-sensitive selection of transcriptional start sites in Escherichia coli. J Bacteriol. 1992;174:514–524. doi: 10.1128/jb.174.2.514-524.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Wassarman KM. 6S RNA: a small RNA regulator of transcription. Curr Opin Microbiol. 2007;10:164–168. doi: 10.1016/j.mib.2007.03.008. [DOI] [PubMed] [Google Scholar]
  • 140.Wassarman KM, Saecker RM. Synthesis-mediated release of a small RNA inhibitor of RNA polymerase. Science. 2006;314:1601–1603. doi: 10.1126/science.1134830. [DOI] [PubMed] [Google Scholar]
  • 141.Cavanagh AT, Klocko AD, Liu X, Wassarman KM. Promoter specificity for 6S RNA regulation of transcription is determined by core promoter sequences and competition for region 4.2 of σ70. Mol Microbiol. 2008;67:1242–1256. doi: 10.1111/j.1365-2958.2008.06117.x. [DOI] [PubMed] [Google Scholar]
  • 142.Cayley S, Record MT., Jr Large changes in cytoplasmic biopolymer concentration with osmolality indicate that macromolecular crowding may regulate protein–DNA interactions and growth rate in osmotically stressed Escherichia coli K-12. J Mol Recognit. 2004;17:488–496. doi: 10.1002/jmr.695. [DOI] [PubMed] [Google Scholar]
  • 143.Gralla JD, Vargas DR. Potassium glutamate as a transcriptional inhibitor during bacterial osmoregulation. EMBO J. 2006;25:1515–1521. doi: 10.1038/sj.emboj.7601041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Gourse RL. Visualization and quantitative analysis of complex formation between E. coli RNA polymerase and an rRNA promoter in vitro. Nucleic Acids Res. 1988;16:9789–9809. doi: 10.1093/nar/16.20.9789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Ohlsen KL, Gralla JD. Interrelated effects of DNA supercoiling, ppGpp, and low salt on melting within the Escherichia coli ribosomal RNA rrnB P1 promoter. Mol Microbiol. 1992;6:2243–2251. doi: 10.1111/j.1365-2958.1992.tb01400.x. [DOI] [PubMed] [Google Scholar]
  • 146.Potrykus K, et al. Antagonistic regulation of Escherichia coli ribosomal RNA rrnB P1 promoter activity by GreA and DksA. J Biol Chem. 2006;281:15238–15248. doi: 10.1074/jbc.M601531200. [DOI] [PubMed] [Google Scholar]
  • 147.Laptenko O, et al. pH-dependent conformational switch activates the inhibitor of transcription elongation. EMBO J. 2006;25:2131–2141. doi: 10.1038/sj.emboj.7601094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Mukhopadhyay J, Sineva E, Knight J, Levy RM, Ebright RH. Antibacterial peptide microcin J25 inhibits transcription by binding within and obstructing the RNA polymerase secondary channel. Mol Cell. 2004;14:739–751. doi: 10.1016/j.molcel.2004.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Nickels BE, Hochschild A. Regulation of RNA polymerase through the secondary channel. Cell. 2004;118:281–284. doi: 10.1016/j.cell.2004.07.021. [DOI] [PubMed] [Google Scholar]

RESOURCES